Rop - MG 2001 PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

Prog. Polym. Sci.

26 (2001) 895±944
www.elsevier.com/locate/ppolysci

Melt rheology of polyole®ns


Markus Gahleitner
Department RAPP, Borealis GmbH, Research Location Linz, St. Peterstr. 25, A-4021 Linz, Austria
Received 8 January 2001; revised 11 March 2001; accepted 12 March 2001

Abstract
Rheology has a key position in polymer research, being an important link in the so-called `chain of knowledge'
reaching from the production of polymers to their end-use properties. A review of the melt rheology of polyole®ns,
which are the most widely used group of thermoplastic polymers today, is given in this paper both in terms of
application and characterisation aspects. The materials are discussed according to their phase structures (single-
and multi-phase polymers) and their chain structures (linear and branched). Aspects of the molar mass distribution,
the chain structure and topology are discussed both from an experimental and theoretical point for the single-phase
systems. For the technically more important types of multi-phase polymers like compounds and blends, the
importance of rheological properties in the development of the phase structure is outlined as well as the possibility
to use rheometry for structure investigations. In any case, the importance of considering the stress or strain history
of a material sample in a rheological investigation is discussed. Finally, an outlook on the present and future
developments in the ®eld of polyole®ns is given. q 2001 Elsevier Science Ltd. All rights reserved.

Keywords: Polyethylene; Polymer blends; Polypropylene; Polymer; Polymer processing; Polymer melts; Review

Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 896
1.1. Material classes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 896
1.2. Rheological properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 898
1.3. Application areas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 899
2. Single phase systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900
2.1. Molar mass effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 900
2.2. Chain structure effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
2.2.1. Linear chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
2.2.2. Branched chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 910
2.2.3. Partially crosslinked chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 919
3. Multiphase systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 920
3.1. Inhomogeneous products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 921

E-mail address: markus.gahleitner@borealisgroup.com (M. Gahleitner).

0079-6700/01/$ - see front matter q 2001 Elsevier Science Ltd. All rights reserved.
PII: S0 0 7 9 - 6 7 0 0 ( 0 1 ) 0 0 01 1 - 9
896 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

3.2. Filled and reinforced polyole®ns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 922


3.3. Crystallizing polyole®ns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 926
3.4. Elastomer blends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 927
3.4.1. Extruder blends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
3.4.2. Reactor blends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 932
3.5. Blends with other polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 935
4. Actual and future trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 938
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940

1. Introduction

For thermoplastic polymers, the knowledge and moreover the design of ¯ow behaviour is essential for
all forms of production and processing, as big parts of these occur in the molten state [1]. Additionally,
rheology has attained a key position in polymer research, being an important link in the correlation chain
(see Fig. 1) from the catalyst over polymerisation and chain structure to processing behaviour and ®nal
properties [2]. Thereby it forms an important knot in the so-called `chain of knowledge' reaching from
the production of polymers to their end-use properties, which has become increasingly important in view
of the increasing speed of material development as a result of quickly changing customer requirements.

1.1. Material classes

Polyole®ns, which are normally de®ned as polymers based on alkene-1 monomers or a-ole®ns, are the
most widely used group of thermoplastic polymers today. Based on their monomeric units and their
chain structures, they can be divided into the following subgroups:

² Ethylene-based materials Ð polyethylenes (PEs) Ð produced under low pressure conditions with
transition metal catalysts of various types and showing a predominantly linear chain structure. This
subgroup includes high density PE (HDPE), medium density PE (MDPE), linear-low density PE
(LLDPE) and other varieties, which are distinguished through the regulation of density and subse-
quently mechanical properties through the incorporation of higher a-ole®ns (mostly butene, hexene
and octene) as comonomers. The linear nature of their polymeric chains can be disturbed twofold: by
longer comonomers like butene, hexene or octene acting as short side chains [3] and by catalysts
forming polymerisationally active oligomers being incorporated further as long chain branches
(LCBs). Examples for the latter case including implications for the rheology of such systems are
given by Yan et al. [4].
² Ethylene-based polymers (PEs) produced in a radical polymerisation under high pressures with
oxygen or peroxides as chain initiators and showing a predominantly branched chain structure.
According to their reduced crystallinities and densities, these materials are termed low density
polyethylenes (LDPEs). A variation of the degree of branching is possible by various measures
like temperature control, peroxide feed, residence time etc., which strongly affects the material's
rheological and terminal properties [5].
² Propylene-based polymers produced with transition-metal catalysts Ð polypropylene (PP) and its
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 897

Fig. 1. Rheological parameters acting as a `link' between molecular structure and ®nal properties of a polymer.

copolymers Ð showing a linear chain structure with stereospeci®c arrangement of the propylene
units. Mostly, the isotactic species Ð iPP Ð is used today, but also syndiotactic Ð sPP Ð or special
stereoblock structures have become technically relevant, as they are available from single-site cata-
lysts. A wide variation of material properties can be achieved with the incorporation of ethylene and/
or higher a-ole®ns in various fashions; single-phase and multiphase materials are possible [6].
² Polymers based predominantly or exclusively on higher a-ole®ns (e.g. poly-butene-1), produced with
transition-metal catalysts and having a linear and stereospeci®c chain structure.
² Ole®nic elastomers based on transition metal or single-site catalysts, with or without the incorpora-
tion of dienes, which make these materials partially crosslinkable. These polymers are normally based
on ethylene and propylene, mostly amorphous with high molar masses and rarely homogeneous in
their phase structures. Nowadays, such elastomers are sometimes substituted with metallocene-based
ultra-low density PEs (ULDPEs) termed plastomers [7], which have a more homogeneous structure
and can be varied in their properties more easily.

Also of importance are an inhomogeneous group of materials that are based inhomogeneous group on
blends of different polyole®ns. The terminology is not completely `sharp' here; extruder-based mixtures
with solid substances (®llers and reinforcements) and elastomers of any kind as well as other polyole®ns are
normally called `compounds', while similarly produced mixtures with non-PO polymers (and also elasto-
mers) as well as reactively produced mixtures incorporating grafting steps are called `blends'. All of these
materials have attained importance in areas where the physical limits of polyole®ns need to be exceeded or
basic characteristics of their chemical nature (e.g. hydrophobicity and apolarity) need to be changed.
Some limiting cases will not be treated within this review. These include wax-like atactic PP (used
e.g. in glue systems and as asphalt modi®ers) as well as ole®n-oligomers produced as constituents for
lubri®cation systems. Common to these materials is their low molar masses, which goes along with
Newtonian behaviour in the molten state or even a liquid nature at room temperature.
898 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

1.2. Rheological properties

Before going into details, the scope of this paper should be clari®ed. It will cover the melt rheology of
polyole®ns, including the limiting case of solidi®cation Ð which is, in case of practically all technically
relevant polyole®ns, crystallisation. From a theoretical point of view, these properties can be split into:

² Linear viscoelastic properties, the range of material behaviour where a linear relationship between
stress and strain exists. These are theoretically the simplest properties, having a direct relation to the
molecular structure or superstructure in case of multiphase systems and being mainly used for
investigation of the same. Normally, the determination of storage and loss moduli (G 0 , G 00 (v ))
combining low material and time demand with high precision [8] is used for this purpose. However,
creep and relaxation measurements Ð J(t) and G(t) Ð have an important position here [9], giving
improved access to the long-time (low frequency/rate) behaviour of viscoelastic materials.
² Non-linear viscoelasticity, indicating the stress (or strain) sensitivity of material behaviour. Techni-
cally most relevant is the classical determination of the ¯ow curve, forming the base for all simple
types of ¯ow modelling. The most simple case of steady-state viscosity measurements is the deter-
mination of the melt ¯ow rate (MFR) as outlined e.g. by Bremner et al. [10], but for meaningful
values a good de®nition of the state of deformation during and even before the measurement is
necessary. Problems in reproducibility and comparability frequently have their origin in a neglect
of these in¯uence factors. Also, a differentiation between stress- and strain-controlled measurements
must be made [11,12]. This has lead to two different philosophies in instrument design with different
strengths and weaknesses, depending on the nature of the investigated systems.

Other non-linear properties include normal stresses (or normal stress coef®cients) and measures for
elasticity. These are mostly relevant in free-surface moulding processes like extrusion, blow moulding or
foaming, and their determination is normally less straightforward. This also puts a practical limitation to
viscoelastic modelling of such processes through the limited availability of the necessary data.
Although capillary measurements have lost some importance in recent years, these instruments are
still very valuable for steady-shear investigations at very high shear rates for the determination of melt
fracture phenomena [13] and related effects (sharkskin structure, oscillations and pumping, spurt effect,
etc.). Mostly, purely optical detection is used for quantifying these effects, which are effectively limiting
the output rates in all extrusion-type conversion processes. No uniform classi®cation of melt fracture
phenomena can be found for different types of polymers, also because of the signi®cant in¯uence of the
polymer chain structure.
Extensional properties still have a special position in the whole ®eld of melt rheology. Partially, the
problem can be de®ned with `what should we measure Ð what is relevant'; and not even for the normal
case of uniaxial extension a standard measuring procedure is available [14,15]. Depending on whether
somebody wants to determine `pure' rheological properties or rather processing behaviour in technol-
ogies with a strong elongational ¯ow component (e.g. ®lm blowing, foaming, coating, etc.) the choice of
an `appropriate' measuring technique will have to be different. Details will be discussed in Section 2.2.2.
Generally, the appropriate range of deformation (strain) or load (stress) will have to be used in
obtaining the relevant parameters for a given process (see Fig. 2). Also, time dependence and history
effects have to be considered.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 899

Fig. 2. Typical viscosity curve of a polyole®n Ð PP-homopolymer, MFR (2308C/2.16 kg) of 8 g/10 min Ð at 2308C with
indication of the shear rate regions of different conversion techniques.

1.3. Application areas

Polyole®ns are nowadays used in practically all application areas of thermoplastics and are processed
with all standard conversion techniques. Despite the fact that they are considered to be `well established'
and `well known' materials, a multitude of new grades continuously appearing on the market and having
partially new structural features are demanding special attention [16±18].
Moreover, the development of polymeric materials can be seen as a `push±pull-approach'. Both the
input of new knowledge and results from the scienti®c side of the development Ð new catalysts, new
polymers or basic facts on ¯ow or solidi®cation processes Ð and the steady change of the market
situation Ð through customer demands, general trends or the overall socio-economic situation Ð are
contributing to the speed of evolution.
The main types of processing, for which rheological requirements need to be considered, are:

² Extrusion Ð ¯at (cast) ®lm, blown ®lm including biaxially oriented ®lm (BOPP), pipe and pro®le
[19,20]
² Extrusion coating and foaming [21]
² Injection moulding [22,23]
² Extrusion blow moulding, injection-stretch blow moulding (ISBM) and thermorming [24]
² Fibre spinning [25]
² Special processes like rotomoulding or powder-slush moulding [26,27]

Processing simulations today are still mostly limited to the viscous part of the behaviour of polymer
melts and frequently make use of generalised newtonian models for description (as shortly discussed in
Section 2.1). This is certainly not suf®cient in case of free-surface or elongational-¯ow dominated
conversion processes. But even if the full viscoelastic nature of a material is taken into account, the
900 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

solidi®cation as an integral part of the forming process still remains open. In case of polyole®ns,
solidi®cation means crystallisation, which will strongly interact with the ¯ow processes as such. This
will be discussed to some extent in Section 3.3.

2. Single phase systems

The case of single-phase (homogeneous) polymers is the most simple one to be considered. Here, the
polydisperse nature of technical systems characterised by the molar mass distribution (MMD) can be
directly correlated to the viscoelastic behaviour. The only additional factor for these systems is the
nature of the polymer chain, its stereostructure and eventual branching, which will strongly affect the
performance.

2.1. Molar mass effects

For a technical polymer, the MMD is a direct result of the statistics of the polymerisation process. In
the process itself, it results from:

² The nature of the catalyst: Conventional Ziegler/Natta type catalysts [6] have a variety of active sites
with different chemical natures and characteristics regarding hydrogen response, comonomer inser-
tion and stereostructure. In contrast, `single-site' like metallocene-based systems [28] have identical
characteristics for each active site, allowing for a much more homogeneous polymer structure. Other
opportunities added by single-site catalysts include the incorporation of a much wider variety of
comonomers (e.g. styrene, conjugated dienes or even polar components) and the unsaturated termi-
nation of polymer chains, the consequences of which will be further outlined below.
² Design and operational mode of the polymerisation reactor: Depending on the reactor geometry Ð
stirred tank, tubular/loop, ¯uzidised bed etc. Ð and the chosen operating conditions Ð temperature,
pressure and ¯ow Ð even identical catalysts will produce different polymers, because of the different
residence time distributions of the system. A totally new range of properties has been opened here just
in recent years by the application of supercritical conditions for ole®n polymerisation in the liquid
phase.
² Monomer composition and feed mode: Even in single-phase systems, differences in catalyst reactiv-
ities not only affect the chain composition, but also the MMD. Moreover, the variation of the catalyst
system (donor type) and the monomer feed composition in serial reactors allows for the control of
tacticity and comonomer distribution over the MMD. The term `bimodal' is frequently used, although
truly bimodal products (i.e. showing two clearly separated peaks in the MMD) are the exception
rather than the rule. Of equal interest and probably less problematic in its consequence to product
homogeneity (see Section 3.1) is the bimodal distribution of tacticity or comonomer content.

Further changes are possible in post-polymerisation processes. In case of PP, degradation with
peroxides in the molten state is normally called `visbreaking' or a `controlled rheology'-(CR)-process
[29,30], resulting in polymers with a signi®cantly more narrow molar mass distributions. An example for
a series of PP-homopolymers produced from one base polymer with different peroxide amounts is given
in Table 1 and Fig. 3. Adjusting the ¯owability (MFR) of a certain grade in the CR-process allows one to
produce a reduced number of reactor grades, thus facilitating production and storage logistics as well as
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 901

Table 1
Degradation series of PP-homopolymers Ð evolution of MMD, rheological and mechanical properties (basic polymer M0 from
standard liquid bulk process, Bis(tert.butylperoxy-isopropyl)benzene Ð DIPP Ð used as peroxide, degradation in COLLIN
50 mm twin-screw extruder at 210±2208C; MMD-data from GPC, MFR ISO 1133 at 2308C/2.16 kg, ¯exural modulus DIN
53452/57, Charpy Impact ISO 179 1 eA Ð V-notch at 1238C; materials as in Ref. [29])

Material cP (wt%) MW (kg mol 21) MW/Mn MFR Flexural mod. Charpy impact
2308C/2.16 kg (MPa) (kJ m 22)
(g/10 min)

M0 0 766 5.5 0.4 1419 7.5


M1 0.026 453 3.5 3.4 1247 4.7
M2 0.051 318 3.1 8.6 1213 3.9
M3 0.108 231 2.8 28 1208 3.0
M4 0.146 181 2.7 51 1175 2.6
M5 0.175 157 2.7 81 1157 2.4
M6 0.24 135 2.5 149 1150 1.9

reducing the `transition' quantities produced when changing the grade in the reactor. CR±PP also shows
a number of speci®c advantages regarding processability and the ®nal material pro®le in mechanics and
optics. Both the type of peroxide used and the design of extruder used (speci®c energy input and
relaxation time distribution) will tend to de®ne the product properties in detail [29].
In contrast, the peroxide treatment of PE-based systems normally results in branching and cross-
linking reactions with an increase of the average molar mass and broadening of the MMD [31]. These
different consequences of radical reactions in the polymer have to be considered when planning or
analysing reactive modi®cation processes [32], which are often involving multiphase systems comprising
both PP-based and PE-based phases. The consequences of this will be further discussed in Section 3.4.2.

Fig. 3. Evolution of the viscosity curve at 2308C (calculated from dynamic test in plate/plate geometry using Cox/Merz-
relation) in the degradation series of Table 1.
902 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Only recently it has been shown that such `buildup'-reactions are also possible and technically
relevant for PP. These radical reactions can either be initiated by gamma radiation under vacuum
[34] or by very high peroxide concentrations [34]. As an example, the development of MFR over
time for different peroxide concentrations are given in Fig. 4. While for low concentrations only
degradation effects are found, the material reacts with chain buildup and viscosity increase at higher
concentrations. Apart from the quantity, also the type and especially the decomposition temperature and
velocity of the applied peroxide have an important effect here. Peroxides with a lower decomposition
temperature becoming active in an earlier stage of the extrusion process are advantageous for buildup
reactions.
Moreover, the absence of oxygen as well as the chain structure and crystalline morphology have been
shown to be essential in the radiation initiation of these reactions [34]. In all cases, long chain branched
structures are created which will be further discussed in Section 2.2.2.
Also for PE, degradation processes are possible [35]. The reaction to free radicals is strongly

Fig. 4. Dependence of the MFR of PP-homopolymer on the decomposition time of a peroxide (Di-tert.butyl-peroxide, DTBP) at
a reaction temperature of 1298C and various concentrations: (1) 4.62; (2) 9.25: (3) 18.5; (4) 37.0; (5) 74.4; (6) 100.0, (7) 139.0
and (8) 200.0 mmol kg 21 PP (from Ref. [34]; reprinted with permission of Wiley, 2000).
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 903

dependent on the catalyst type. For chromium-based catalysts, branching is normally predominant, while
for titanium-based catalysts, degradation is more important.
As a result of all these possibilities, the shape and moments of the MMD can be varied over a wide
range. In principle, correlations to all rheological properties are possible from the MMD if all chains are
fully linear. Based on a spectral assumption, for a polydisperse polymer, correlations can be setup
between the molar mass of each fraction and the respective relaxation time, but interactions between
the different molar mass fractions have to be considered as well. The mathematical formulations
developed for this are called `mixing rules' [1].
As an example, one can start with the formulation for Lodge's `rubberlike liquid' Ð model relating
the stress and strain of a viscoelastic substance
Zt
s…t† ˆ 2p1 1 m…t 2 t 0 †C 21 …t 0 † dt 0 …1†
21

where s stands for the stress in the system, m for the memory function, C 21 for the Finger tensor
describing the state of deformation and the term -p 1 for the isotropic pressure contribution. If the
memory function is developed in the usual way of a generalised Maxwell model (with N elements, where
N should allow a density of 1±2 relaxation times per decade), one obtains
X
N
m…t 2 t 0 † ˆ gi exp ‰2…t 2 t 0 †=ti Š …2†
iˆ1

where gi, t i are the N pairs of relaxation times and strengths of a discrete relaxation time spectrum
(RTS). These are connected to the relaxation modulus in the following way:
X
N
G…t† ˆ gi exp…2t=ti † …3†
iˆ1

The most simple correlation between rheology and the molar mass distribution (MMD) can be
formulated between MW and the zero shear viscosity h 0 (also valid for the viscosity in dilute solution).
Two regions are clearly separated: below the critical molar mass MC a relation of h 0 / MW holds, while
3;4
above MC the proportionality changes to h0 / MW : The critical molar mass MC < 2ME, where ME is the
average molar mass between two entanglements in the system. Based on this it is also possible to
establish a more profound relation between rheology and MMD. Using the above discrete relaxation
time spectrum, the relaxation times can be converted by
ti ˆ aMi3;4 …4†
to the average molar masses of a respective number of fractions of the MMD. A closer look however
shows the necessity to take the in¯uence of the environment on every molecule into account; the relation
then changes to
ti ; m ˆ aMi3;42b MW;m
b
with b < 1; 4 …5†
Formulations like Eq. (4) are the basis of the so-called mixing rules for the realistic description of
polydisperse systems.
In recent years, a multitude of research works has been published dealing with the problem of
904 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

interconversion between linear viscoelastic data and the MMD of a polymer. Roughly, these papers can
be split into three categories:

² Development of empirical or semi-theoretical correlations between singular parameters of the linear


viscoelasticity Ð like zero shear viscosity h 0, equilibrium compliance Je,0, crossover frequency and
modulus of G 0 ,G 00 (v ), etc. Ð and parameters of the MMD like weight average MW or polydispersity
MW/MN [36±42]. In some cases, also more general rheological data like the melt ¯ow rate (MFR) are
included within these considerations [10]. These correlations have a general problem of limited
validity and are restricted to narrow groups of materials only.
² Full-scale interconversions between the MMD and linear viscoelastic curves like G(t) or G 0 ,G 00 (v ),
which make up the majority of work in this area [43±51]. These correlations can also be expanded to
predictions or interpretations of the processing behaviour [25]. A common problem of this second
type is the fact that the calculation is straightforward only in one direction, i.e. from the MMD to
viscoelasticity. The reverse is an ill-posed problem which allows for the calculation of different forms
of MMDs based on one set of linear viscoelasticity data.
² Another independent way of interrelating MMD and rheology data is the application of statistical
methods like multivariate analysis (MVA). With MVA, big data sets like complete MMD- and
rheology-curves can be easily interrelated, creating a set of `latent variables'. Although the resulting
equations and related parameters are not necessarily physically meaningful, this method is very well
suited both in quality control and for the determination of development trends [52,53]. Moreover,
other parameters Ð of a mechanical or an analytical nature Ð can be included in the correlation.

The most `popular' parameter correlations are the ones involving the zero shear viscosity, which is
theoretically a nice parameter, but dif®cult to determine in case of wide molar mass distributions. An
example from earlier work within our group [42] is given in Fig. 5 for two different series of
PP-homopolymers from a commercial process and from a pilot-scale polymerisation, based on 4th

Fig. 5. Correlation between weight average of the MMD and zero shear viscosity for three series of PP-homopolymers of
different origin (K Ð Ziegler/Natta-(ZN)-type catalyst, CR-grade; £ Ð ZN-type catalyst, reactor grade; V Ð Single-site
catalyst); data from Ref. [42].
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 905

generation Ziegler/Natta type (ZN) or metallocene catalysts (MC). The zero shear viscosity h 0 was
determined from creep measurements, which are advantageous for such measurements in the long time/
low frequency range as has also been demonstrated [54,55]. Other frequently used empirical relations
involve the `crossover'-parameters of the dynamic moduli
GC ˆ G 0 …vC † ˆ G 00 …vC † and vC …6†
which have no clear theoretical basis but are easier to determine. According to Zeichner and Patel [36],
v C can be correlated to MW and 1/GC to the polydispersity MW/MN. The validity of this correlation for
materials with a similar shape of the MMD has been demonstrated several times. In both cases it is easier
to predict the average molar mass than the polydispersity.
As a consequence, a number of other parameters have been de®ned over time Ð mostly by industrial
rheologists Ð which are related to the polydispersity. A nice review of several of these parameters has
been given by Steeman [40], who used numerically calculated rheological data based on `model-MMDs'
to evaluate the applicability and limits of such quantities. While the correlations between MW and h 0 of
the general form
a
h0 ˆ hK MW with a ˆ 3:2 to 3:6 …7†
were found to be hardly affected by the polydispersity (MW/MN) as well as by higher moments of the
MMD, no such simple relation could be established for Je,0. Different formulations like
2 MZ11 MZ
Je;0 ˆ …8†
5 MW rRT
or
 a
MZ
Je;0 / …9†
MW
were tested. Examples of the results are presented in Figs. 6 and 7, where the asymmetry parameter
H ˆ …MZ =MW †=…MW =MN † …10†
was varied and a double-reptation model combined with BSW-parameters [45] was used for calculating
the linear-viscoelastic material functions from the simulated MMD. As can be seen, none of the applied
formulations allows for a reasonable prediction of the Je,0 value. The same applied for empirical
parameters like the MODSEP-parameter [38], which is calculated from the distance between G 0 and
G 00 on the frequency axis like
MODSEP ˆ v…G 0 ˆ 10:000 Pa†=v…G 00 ˆ 10:000 Pa† …11†
Generally no uniform correlation to the MMD polydispersity could be established. This is also true for
the `shear thinning index' (SHI), which was developed in Borealis R and D and is used within our
organisation as a quality control parameter. The SHI-value can be calculated according to
SHI…0=A† ˆ h0 =h…Gp ˆ A kPa† …12†
where the value of A can be 10, 50 or 130 depending on the molar mass area of the polymer in question.
Another possible way to determine MMD-related parameters is the application of generalised
Newtonian ¯ow models for the viscosity curve h (g 0 ). One example is the generalised Carreau or
906 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 6. Steady-state compliance of model polymers plotted as a function of Mz11Mz/MWMn; the straight line indicates the
dependence proposed by Ferry (from Ref. [40]; reprinted with permission of Steinkopff Verlag, 2000).

Carreau±Yasuda model [37] of the form

h…g 0 † ˆ h0 ‰1 1 …lg†BŠ…n 2 1†=B …13†

where h 0 is the zero shear viscosity, l the characteristic relaxation time, n the power law Ð index, and B
is the transition parameter for which both h 0 and l can be correlated to MW, while B can be correlated to
the polydispersity [37].

Fig. 7. Steady-state compliance of model polymers plotted as a function of Mz/MW; the solid curve indicates the dependence
proposed by Kurata (from Ref. [40]; reprinted with permission of Steinkopff Verlag, 2000).
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 907

Fig. 8. Dynamic moduli of a PP-homopolymer (reactor grade; MW ˆ 165 kg mol21 ; MN ˆ 25:1 kg mol21 ; MW =MN ˆ 6:6) at
1858C from experiments Ð A storage and W loss modulus Ð and calculated from full spectrum (full line) and reduced
spectrum (dashed line) using the double reptation concept (from Ref. [49]; reprinted with permission of Wiley, 2000).

The present status of full-scale interconversions is somewhat hard to judge, as the number of papers
and even theories proposed is in contrast to the efforts taken to experimentally verify the theories. An
example for the possibility to calculate linear-viscoelastic data from the MMD Ð taken from a work by
Carrot et al. [49]Ð is given in Fig. 8. In this paper, a series of 10 different polypropylenes covering a
certain range of MW and polydispersity was investigated both rheologically and with GPC, combining
the obtained data with the double reptation model, for which the ®rst part of the relaxation modulus of a
polydisperse sample composed of N types of different species with relaxation times t i and corresponding
weight fractions Wi can be expressed as
XX
GN …t† ˆ G0N Wi Wj F 1=2 …t; ti †F 1=2 …t; tj † …14†
i j

in case that both Mi and Mj are higher than MC. In this case, the running index i denotes the relaxing chain
and j the environment; it represents the effective contribution of chain entanglements. F is a function
allowing to reach GN …t† ˆ G0N (plateau modulus) for small values of t. The part of the polymer with
molar mass below MC can additionally be represented by a Rouse spectrum such as
" #
p2 0
X Wi
GR …t† ˆ MG F…t; ti † …15†
6 e N i Mi

In this case, the relaxation time of the fraction remains unchanged by the environment. The third part
of the system are the chains between entanglements and the can also be formulated in a Rouse spectrum
as
" #
p2 0
X
GE …t† ˆ MG W F…t; tee † …16†
6 e N i i

where t ee is the theoretical relaxation time of a chain of length Me. The three contributions can then be
908 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

summed up to give
G…t† ˆ GN …t† 1 GE …t† 1 GR …t† …17†
Based on these equations and some basic data for the polymer Ð namely Me and G0N Ð the MMD as
determined by GPC can be converted to a relaxation spectrum with a high number of modes and, ®nally,
into linear viscoelastic material functions like G(t) or G 0 ,G 00 (v ). As Carrot et al. [49] shows, this can be
greatly facilitated by reducing the number of modes used for the actual conversion without losing too
much information. The actual comparison between calculation and measurement in Fig. 8 shows,
however, the problem of a conversion in this direction as a deviation in the low frequency region.
This is caused by the limited sensitivity of GPC at high molar masses. Consequently, the precision of
these calculations tends to increase with reduced MW and polydispersities.
In any case, most correlations are presently limited to shear ¯ow. It has however also been demon-
strated that extremely wide or bimodal MMDs are also re¯ected in the behaviour under extensional ¯ow
[19,56]. Strain hardening effects Ð which will be discussed in Section 2.2.2 Ð can appear for such
materials, signi®cantly improving their processability for example in ®lm blowing [57] and blow
molding [58]. Non-linear formulations like varietes of the K-BKZ form
Zt
s…t† ˆ 2p1 1 m…t 2 t 0 †h…l1 ; l2 †C 21 …t 0 † dt 0 …18†
21

with p 1 again denoting the isotropic pressure contribution, m the purely time-dependent memory
function, h the damping function depending on deformation and deformation rate and C21 the Finger
tensor used as a deformation measure, have to be used here. In addition to determining the time
dependent component of the material behaviour (m(t 2 t 0 ) or g(t ), relaxation time spectrum), the
deformation dependent component (h, damping function) also has to be determined.
In addition the continuum approach needs to be replaced by a type of element modelling like the
FEM-calculation, taking into account the various strain histories of different segments of the moving
melt.
Apart from the purely rheological and processing behaviour [59], changes in the MMD always have
some `side effects', which are of relevance for the mechanical and optical properties of the polymers.
Examples for the case of PE are given by Shroff et al. [60] and by Jordens et al. [41], for PP by
Tzoganakis et al. [61] and by Bailey and Varrall [62].
The concept of `tie molecules' is essential to the work of Huang and Brown [63], who demonstrated a
signi®cant improvement in the slow crack growth dependence on the MMD shape and especially on the
presence of high molar mass fractions. For linear PE grades with very high toughness requirements,
however, the amount and distribution of comonomer plays an even bigger role for the ®nal properties
[64]. Single-site catalysts allow for an improved design of polymer properties by controlling the
comonomer incorporation over the MMD for PE or the tacticity distribution for PP.
Another good example is the effect of the MMD on the mechanics for PP-homopolymers, which can
be explained by differences in the crystallisation behaviour in¯uenced in turn by the MMD. An increase
in the polydispersity does not only lead to an increased nucleation density in quiescent crystallisation
[65], but also to enhancement of strain-induced structure development [66,67]. Details of this effect will
be discussed in Section 3.3. It must however be clearly stated that such correlations are practically never
universal. MC-based PP-grades for example, which have been shown to require signi®cantly different
machine settings in conversion processes [68] do not ®t into this picture because of their reduced
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 909

Fig. 9. In¯uence of the polydispersity MW/MN of PP-homopolymers based on Ziegler/Natta type catalysts on the nucleation
density Nc at 1108C showing the nucleating effect of high molar mass fractions; data from Ref. [69].

tacticities and the fact that MMD-measurements do not necessarily account for very small fractions of
high molar masses [69]. The general dependence of the nucleation density on the polydispersity is
presented in Fig. 9, from which the effect of long-chain molecules acting as nuclei on crystallisation
is re¯ected in an increase of NC with MW/MN.
From both examples it becomes clear that polyole®ns with a wide or even bimodal MMD can be
advantageous with respect to both processing behaviour and end-use properties, leading to important
developments in this direction [70]. A possible problem in the technical production of such materials is,
however, inhomogeneity induced by the presence of high- and low-viscosity fractions. This will be
discussed in Section 3.1.

2.2. Chain structure effects

A closer look at the structures of polymers shows us that there are actually two levels or dimensions of
`structure'. For a given polymer, the local chain structure can be affected by the stereochemistry Ð
wherever it is relevant, e.g. for PP and PB-1 Ð and by comonomers forming zones of higher ¯exibility
in the chain Ð like ethylene in PP Ð or very short side chains Ð like butene in PE. The global structure
or topology, which affects rheological behaviour to a greater extent, is determined mainly by chain
branching. Normally, chains with local structural disturbances are still considered to be linear, although
their viscoelastic behaviours may be signi®cantly altered.

2.2.1. Linear chains


Copolymerisation is used to control the mechanical and optical properties of polyole®ns for matching
customers' requirements. In the case of a purely random incorporation and distribution of the comono-
mer, (which is mainly achieved at low concentrations) the system remains single-phase. Apart from the
amount of comonomer, the catalyst and process type will also determine the transition from this simple
case to the more complex case of heterophasic copolymers which will be discussed later in Section 3.4.2.
For polyethlyene, where stereochemistry does not play any role, the local structure is only affected by
910 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 10. Zero shear viscosity h 0 versus weight average molar mass MW in a log/log-plot for syndiotactic PP (A) and isotactic PP
(X); note the identical slope but different position of the line ®ts (from Ref. [77]; reprinted with permission of VCH Wiley,
2000).

copolymers acting as very short chain branches. The effect of comonomers on rheology in case of
LLDPE-type polymers has been investigated several times [19,62,71]. It appears that while the como-
nomer has a certain effect on rheology Ð especially on the relation between MMD and viscoelasticity
Ð signi®cant changes like the appearance of strain hardening are only found in case of long side chains
(.MC), which appear frequently with the use of single-site catalysts [4,72] and are even favored on
purpose to improve the processability of such materials [73]. Such materials are directed at application
areas presently covered by LLDPE/LDPE-blends [74,75] because of melt strength requirements, e.g. in
the blow molding of large containers or ®lm blowing.
For polymers with stereospeci®ty, tacticity effects have to be considered alongside with comonomer
insertions. In case of isotactic, atactic and syndiotactic PP produced with metallocene catalysts, separate
relations between MW and h 0 as well as other pairs of MMD- and rheology-parameters were found for
different degrees of tacticity by Friedrich [18,76,77]. The measurements, which were carried out in a
combination of creep and dynamic tests to cover a wide time/frequency-range, showed changes in the
activation energies (E1a-values were found to be 31 kJ mol 21 for iPP and 57 kJ mol 21 for MC-based
sPP), the h 0/MW-relation (see Fig. 10) the plateau modulus and even the entanglement molar mass (ME).
An explanation can be given based on the Rouse model. If sPP has a different and more `bulky' chain
conformation (predominant all-trans) in the melt, this will change both the effective tube diameter and
friction factor of the system, leading to a higher relaxation time for molecules with identical length.

2.2.2. Branched chains


Polymers with long-chain branching (LCB polymers) exhibit a completely different rheological
behaviour, which has only recently been formulated for the ®rst time in a consistent viscoelastic
model. The `pom±pom'-model developed by Mc Leish and Larson [78] is based on a phased reptation
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 911

Fig. 11. Transient uniaxial extensional, planar extensional and shear viscosity of an 11-mode pom-pom melt in start-up plotted
against measurement results for `IUPAC X' LDPE, shear/elongation rate 0.01 s 21 and temperature 1408C (from Ref. [79],
reprinted with permission of The Society of Rheology Inc., 2000).

process of the side- and backbone-chains of the system and has meanwhile also been applied for the
description of technical LDPE [79,80].
The assumption of this model is a mixture of pom±pom molecules consisting of a backbone and a
number of q . 1 dangling arms at each end (for q ˆ 2; an H-structure is reached). A system consisting of
such molecules will have two timescales of relaxation, the shorter one resulting from the relaxation of
the dangling arms alone having an effective relaxation time of
" #
15 …1 2 x†2 …1 2 x†3
ta …x† ˆ t0 exp s 2 …1 2 fb † with Sa ˆ Ma =Me …19†
4 a 2 3
where t 0 is the relaxation time for a fully retracted arm, x is the relaxing arm fraction, Ma the molar mass
of each arm and f b the backbone fraction of the system. The second, longer contribution results from the
relaxation of the backbone part, for which the relaxation time can be calculated as
4 2
tb ˆ s f t …0†q with Sb ˆ Mb =Me …20†
p2 b b a
where Mb is the molar mass of the backbone. Based on this, a constitutive equation for a polymer
consisting of such molecules can be constructed [78].
In a recent paper by Inkson et al. [80], the model performance has been checked against the behaviour
of different LDPE types, where data for shear and extensional viscosity were taken into account. An
example including three different deformation types is given in Fig. 11. The good accordance between
model calculations and measured results was achieved for a very limited number of `modes', i.e.
different molecular species.
Apart from a change in the viscosity curve and the linear-viscoelastic behaviour in shear, which is
schematically presented in Fig. 12, the material's behaviour in extensional ¯ow is altered completely.
In a transient experiment, strain hardening effects are observed resulting from an increased resistance of
912 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 12. Schematic presentation of the change of the viscosity curve and the storage modulus with long chain branching; linear
polymer Ð solid lines, LCB polymer Ð dashed lines.

the material to disentanglement. For some types of conversion such as ®lm blowing, the blow moulding
of large containers, foaming etc. this is of great importance for processing behaviour. The extent and
strain dependence of this effect is determined by both composition and branching distribution effects
[5,75,81,82].
Moreover, LCB polymers normally show a distinct dependence on their ¯ow histories. Technically
this is used for the shear modi®cation of LDPE, but it can also occur as a processing problem, for
example in the foaming of PP with long chain branching (LCB-PP). Various model assumptions have
been made for this process which is especially characterised by its reversibility and its drastic conse-
quences on the processing behaviour [83±85].
An example of the effect is given in Fig. 13 for a modi®ed polypropylene with long-chain branches.
These investigations were initiated in the course of developing PP with high melt strength (HMS-PP) for
determining the effect of different processing conditions on the material performance. For this purpose,
the Rheotens-setup, which will be described in some detail below, was combined with a feed extruder
and a melt pump including a bypass valve, allowing the maintenance of a constant die velocity while
varying the shear energy input in the extruder. This combination had originally been proposed by
Wagner for allowing a rheologically meaningful evaluation of Rheotens results.
The materials were extruded at extrusion pressures between 20 and 300 bar; the strands coming out of
the die of the extruder were collected for the further investigations. Rheological investigations done
before and after the extrusion were: Rheotens experiments at an acceleration of 120 mm s 22, creep and
oscillatory experiments. Additionally, the molar mass distribution was examined before and after the
extrusion by GPC in order to see whether mechanical scission of chains occurred in the extruder. As Fig.
13 shows for the case of an LCB-PP, the melt strength decreases dramatically with the extrusion
pressure, while drawability remains rather constant. While no signi®cant effect of this treatment
could be seen in the dynamic moduli, the creep measurement resulted in a change of the equilibrium
compliance Je,0 which proved to be fully reversible upon dissolution in hot xylene and solvent evaporation.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 913

Fig. 13. Effect of pre-shearing in the feed extruder on the melt strength (Rheotens curve) of a chemically modi®ed PP with
long-chain branching (n ˆ 0:15 per 1000 C).

The changes in rheological behaviour caused by extrusion occur without any change in the molar mass
distribution; that means that no chains are cut by this process. Similar effects were found for LDPE,
while no changes appeared for a linear PP and only marginal changes for a bimodal HDPE.
The changes induced by this `shear modi®cation' can also be seen in the die swell, as shown in Fig. 14.
For this investigation by Rokudai [83], a technical LDPE sample (over-stabilised with antioxidant) was
sheared intensely in a Brabender twin-blade kneader for different times and then tested in a capillary
rheometre recording especially the die swell. Performing this operation at a lower melt temperature reduced
the effect; by dissolution in hot xylene and evaporation of the solvent the effect could be reverted completely.
Technically, this process can be used for improving processing behaviour [86] and product quality, espe-
cially surface smoothness and transparency. The reduction of the melt elasticity leads to a higher critical
shear stress and less melt fracture, allowing for an output increase in conversion processes.
In general, various processing effects are induced by LCB. For example, the reduction of melt fracture

Fig. 14. Effect of shearing in a twin-blade kneader on the melt elasticity (die swell) of an LDPE melt (MFR 1908C/5 kg ˆ 8.1 g/
10 min; density 0.914 g cm 23); 0 Ð Brabender shearing at 1308C, D Ð Brabender shearing at 1908C, ®lled symbols represent
materials after solvent treatment (from Ref. [83]; reprinted with permission of Wiley, 2000).
914 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Table 2
Shift factors for time/temperature superposition determined from dynamic moduli measurements in plate/plate-geometry and
activation energies calculated from an Arrhenius plot (LCB-PP Ð Daploy HMS 130D, LDPE Ð Daplen 1840 D, HDPE Ð
Daplen BF 5272 bimod, PP Ð Daplen BE 50)

LCB-PP LDPE HDPE PP

T (8C) aT T (8C) aT T (8C) aT T (8C) aT

260 0.484 210 0.676 210 0.890 240 0.657


230 1.000 200 1.000 200 1.000 220 1.000
200 2.155 190 1.480 190 1.191 200 1.504
Ea ˆ 52.2 kJ mol 21 Ea ˆ 72.9 kJ mol 21 Ea ˆ 27.1 kJ mol 21 Ea ˆ 41.8 kJ mol 21

[2] by addition of LCB fractions to otherwise linear polymers has been shown for PE [87] and PP [13];
the effects are however not identical for both polymers. Apart from the fact that the MMD plays an
identically important role in de®ning the sensitivity of a certain material to melt fracture as the degree of
branching, the relation is not a linear one. While a certain amount of melt elasticity (secured by MW/
polydispersity and/or branching) appears to be required for stable extrusion, too much melt elasticity
becomes detrimental again. This has initiated attempts like the use of hyperbranched processing aids
[88] to boost the upper limit of processing speed in ®lm extrusion. Such rather expensive solutions will
however have to compete with other technical approaches like die coating or the addition of ¯uorinated
polymers as processing aids (at least the latter one is a well-established method in the case of HDPE).
Also the behaviour in ®bre spinning [89] and the stability in ®lm blowing [5] can be improved, both
due to a strong elongational ¯ow component in the process. In case of PP, LCB also leads to a signi®cant
increase of the nucleation density and, consequently, to a higher crystallisation temperature of the
materials.
LDPE can be considered as the `classical' LCB material because of its long and widespread applica-
tion for a variety of purposes. A good deal of the popularity of this material results from its uncompli-
cated behaviour in ®lm blowing which is a result of its molecular structure. Due to the purely statistical
creation of side chains in the high-pressure radical polymerisation process, it has a very complex
structure and the problem of a well-de®ned production as well as the characterisation of the resulting
product has been addressed frequently in recent years [90±92]. Branching effects are in principle also
recognised in shear ¯ow [93], where LCBs lead to an increased slope of the viscosity curve at compar-
able MMDs (see Fig. 12). Such measurements and also conventional dynamic rheometry can be used for
the determination of the activation energy from the zero shear viscosity or shift factor values, demon-
strating a signi®cantly higher temperature dependence of the viscoelastic material functions for
branched systems. As Table 2, in which shift factors (aT) and activation energies calculated from an
Arrhenius-type temperature dependence

aT ˆ ek…T2T0 † with k ˆ Ea =R …21†

clearly shows, this is equally true for PE and PP.


As mentioned before, LLDPE-grades based on metallocene (MC)- or single-site-catalysts can also
show a certain extent of LCB. This is a side-effect of MC catalysts, which are able to incorporate vinyl
terminated chains as comonomers. In industry, this has been recognised as a possible way to improve the
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 915

Fig. 15. Time dependence of the elongational viscosity in uniaxial extension at 1508C for two single-site based LLDPEs (see
text for explanation) in comparison to a conventional LDPE (materials as in Ref. [95]).

processing behaviour of these narrow MMD materials [5,73,94,95]. The catalyst system used as well as
the polymerisation conditions play an important role in the amount of LCB reached in the ®nal product.
An example comparing two different MC-LLDPEs with and witout LCBs as a result of the used
catalyst to LDPE is given in Fig. 15. In a comparative study [72,95] several different single-site
(metallocene) catalysts were used for producing LLDPEs with hexene-1 as a comonomer. While the
actual comonomer content dominating the density of the resulting materials was found to have very little
effect on the rheology, the catalyst type determined whether predominantly linear or branched structures
were produced. In Fig. 15, the measuring results of a MuÈndstedt-setup (see below for explanation) are
given for one product each of [n-but-Cp]2ZrCl2 (Cat. 2, NP5) and a siloxy-substituted derivative of
Et(Ind)2ZrCl2 (Cat. 4, NP1), both in combination with MAO. In addition to the strain hardening effect
found for some of the products, the branched structure of these materials was also proven by determina-
tion of the activation energy and in combined GPC/viscometry measurements.
In contrast to PE, there is presently no published way to produce PP with a branched structure directly
in the reactor. LCB- or high melt strength Ð PP (HMS-PP) is therefore produced by post-reactor
modi®cation with various techniques [96,97]. Technical processes today combine the creation of radi-
cals by irradiation or the addition of peroxides with a local crosslinking step, with our without the
application of a bifunctional crosslinking agent. The absence of oxygen is very important in this process,
as has been shown in other circumstances [33,34] in Section 2.1.
As the formation of gels must be avoided, the achieved degrees of branching are signi®cantly smaller
compared to LDPE, which is already obvious from the differences in Ea between the respective linear
and branched materials in Table 2. The altered rheological properties of this material [98,99] lead to
916 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

important changes in the processing behaviour. While foam production was originally the main target of
this development [100], the possibilities of melt strength improvement, also in mixtures of linear and
LCB-PP are meanwhile being recognised in other areas as well [101,102]. An example comparing
differently produced HMS-PP grades is given in Fig. 16; the three differently modi®ed materials are
based on the same linear PP-homopolymer [97]. According to this study by Sugimoto et al., the
application of ionising radiation and peroxide treatment leads to different branching structures, which
however exhibit very similar strain hardening behaviours. Generally, branching occurs mostly in the
high molar mass fractions of the polymer, with the potential danger of the formation of crosslinked
particles (permanent gels).
In contrast to linear polymers, where linear-viscoelastic techniques are a well established experimen-
tal base for rheological characterisation, this cannot be said yet for branched systems. Some of the most
important rheological characterisation techniques for LCB will therefore be outlined here below:
X Rheotens-test. This setup copies industrial spinning and extrusion processes. In principle (see Fig.
17) a melt is pressed or extruded through a round die and the resulting thread is hauled off. The stress on
the extrudate is recorded as a function of melt properties and measuring parameters (especially the ratio
between output and haul-off speed, practically a measure for the extension rate). The method can be used
for a wide variety of materials, but it has some principal problems, which may be separated into two
categories: (a) ¯ow geometry of the spinning process, and (b) problems of thermal equilibrium.
Concerning (a) it has to be said that the spinning experiment is a non-stationary test. Even if for each

Fig. 16. Time dependence of the elongational viscosity in uniaxial extension at 1808C for an unmodi®ed (linear) PP (D) and
three differently modi®ed LCB-PPs produced by ionising radiation (B, C) and peroxide addition (E) with different branching
degrees (from Ref. [97]; reprinted with permission of Wiley, 2000).
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 917

Fig. 17. Schematic presentation of the Rheotens setup (left) and the Meissner setup in the original form (right) for determination
of elongational properties of polymer melts.

extension rate a steady state with constant shape of the extrudate can be reached, the ¯uid elements are
not necessarily in equilibrium as a consequence of their deformation prehistories. As the effect of
molecular orientations plays a major role on rheology, the measurement is not only in¯uenced by the
intentional parameters Ð die diameter, output rate haul-off speed Ð but also by the prehistory of the
melt. With this, the results of the measurement

e 0 ˆ …2pvRw =H† ln …8pv Rw =DR† …22†

and

sE ˆ l…F=pR2 † ˆ …v8Rw F=R3 D† …23†

where R is the die diameter, H the distance between die exit and takeup point (extrudate length), RW the
diameter of takeup-wheels, v the rpm of takeup-wheels D the apparent (wall) shear rate in the die, and F
is the takeup-force (extrudate stress) can be used for calculating hE ˆ sE =e 0 only within limits [103]. A
time dependence of the quantities can hardly be investigated. Moreover one obtains average values
across the whole length of extrudate, in which the extension rate and therefore also the extensional
viscosity vary over a wide range. Regarding (b) it is clear that this is no isothermal experiment, and that
only for the case of high haul-off speed is extrudate at nearly the same temperature as the material in the
reservoir (where the measuring temperature is normally determined. This problem of thermal equili-
brium is slightly reduced by weak heat transfer between the surrounding air and the polymer melt. In
semicrystalline polymers it must furthermore be considered that high overall extensions accelerate the
crystallisation process, thereby in¯uencing the mechanical properties of the `melt'.
Details about a rheologically meaningful evaluation of this test, which have been developed in
wagners group can be found in the literature [104]. The combination of single run results obtained at
different die geometries into so-called `grand master curves' allows for the calculation of apparent
918 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 18. Apparent elongational viscosity h (g 0 ) of LDPE A 18 as calculated from Rheotens mastercurves for different die exit
velocities v0 at T ˆ 1908C; comparison to shear velocity h (g 0 ) (from Ref. [104], reprinted with permission of The Society of
Rheology Inc., 2000).

elongational viscosities. A curve example is given in Fig. 18. Note the dependence of the die exit
velocity, which is in accordance with the pre-shear effects discussed before.
X Meissner-setup [105,106]. In this setup, one tries to approach the ideal assumption of an in®nite
sample extended at constant rate while overcoming the problem of sample ®xation with ®xed clamps,
which can cause local damage and premature break of the sample in the clamping region. This problem
was ®nally solved in the mobile clamping device developed by Meissner (see also Fig. 17). The sample
¯oats on an oil bath (silicone oil), which also facilitates temperature adjustment, but may create
problems through oil absorbed by the sample, thus changing its properties. In the latest development
of Meissner's apparatus the bath was therefore substituted by a cushion of nitrogen gas emerging from a
porous surface.
Even here, one problem remains: the relation between extensional rate and speed of extension. As the
sample thickness decreases, reaching e 0 ˆ const: requires a logarithmically increasing speed of exten-
sion. This makes a wide RPM-range of the drive elements necessary, even more complicated in connec-
tion with suspension elements which must allow a precise stress measurement. All in all, high values of
total extension and especially stationary terminal values appear to be hard to reach. Great care must
furthermore be taken in sample preparation as only a homogeneous and isotropic samples can guarantee
representative results.
X Laun/MuÈndstedt-setup [9,107]. While using the same sample geometry as the Meissner setup, the
®xation problem is overcome here by pasting the sample ends to ¯at carrier plates avoiding any damage
in this way. The geometry is operated with verital samples, reducing suspension problems but also
limiting the applicable temperature range. As a possible way around the speed dilemma mentioned
before, creep-type measurements (constant stress applied) can be used. In thus way, stationarity of the
extension test is achieved much faster.
X Converging ¯ow analysis [14,15]. In this technique, which is also called the `Cogswell-method', the
pressure drop in the in¯ow zone of a capillary rheometre is used for calculating extensional properties of
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 919

Fig. 19. Stationary viscosity in shear and extension for PP homopolymer (Daplen BE 50, MFR 2308C/2.16 kg ˆ 0.4 g/10 min)
at 2308C (h calculated from dynamical data, h E from stagnation point Ð apparatus as described in detail in Ref. [92]).

polymer melts. Assumptions of the shear/extension rate dependence of shear and extensional viscosity
as well as ®rst normal stress coef®cient C 1 are required for being able to do this calculation and the
geometry of the entry zone is important. The comparability to other methods is not good [14]; the results
are only suitable for comparing similar materials.
X Opposed jet setup [92]. This measuring geometry was originally developed for low viscosity ¯uids
like spinning solutions and only later adapted for polymer melts. In this geometry, two deformation
modes are possible: With a peripheral in¯ow and out¯ow through a pair of opposed dies, a uniaxial
extensional ¯ow in the stagnation point in the centre is reached, while with an in¯ow through the dies a
biaxial extensional ¯ow in the midplane vertical to the die axis develops. Apart from the constructive
solution, the development of a mathematical procedure for evaluation was decisive. The main disad-
vantage of the system is the fact that in principle only steady-state measurements are possible. Startup
effects can be detected only approximately. As an example, a comparison of steady state shear- and
extensional viscosity for a PP melt (MFR (2308C/2.16 kg) ˆ 0.4 g/10 min) is presented in Fig. 19.
X Apart from these measurements of extensional viscosity, LCB-structures can also be investigated
indirectly via linear region measurements in oscillation, relaxation or creep [108,109].
In any case, the results can be correlated to processability, mainly for processes with a high elonga-
tional component in ¯ow like ®lm blowing [106], foaming [100], ®bre spinning or paper coating, but
also for the secondary stage of blow molding and ISBM and special cases of injection molding.

2.2.3. Partially crosslinked chains


As mentioned before (see Section 2.1), radical reactions are frequently used in the post-polymerisa-
tion modi®cation of polyole®ns. Partial crosslinking of a thermoplastic material can be technically
interesting for various reasons, from creating melt strength for processing behaviour over reaching
rubberlike elasticity to an improvement of the heat de¯ection temperature.
In principle, the changes of rheological behaviour with the degree of crosslinking can be seen
completely only in a parallel measurement of viscous and elastic parameters of the system. A decisive
point in the network evolution is the gel point [110], from which onward the material shows predominantly
920 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 20. Schematic plot of steady shear viscosity h and equilibrium modulus G1 of a crosslinking polymer, tc indicates the time
of the gel point (based on information from Ref. [110]).

solid-like behaviour (see Fig. 20). For materials where thermoplastic processing is required, higher
degrees of crosslinking are not relevant.
An example in the area of polyole®ns where high crosslinking densities are of interest are EPR- and
EPDM-elastomers. Elastomers normally have their own kinds of rheologies, especially if the materials
have a further tendency to react and crosslink in the molten phase. A frequently used quantity to de®ne
the viscosity of elastomeric materials is the Mooney-viscosity which is determined in a plate/plate setup
at a given time after heating up from room temperature and therefore combines melting and possible
crosslinking effects in one result.
These materials are hardly ever used in pure form; much more relevant is the behaviour in hetero-
phasic systems [111±113], which will be treated in more detail in Section 3.4. An example for the effect
of a partially crosslinked EPDM dispersed in a PP matrix is given in Fig. 21, where the reduction of the
in¯uence of the disperse phase on the system rheology with increasing shear stress (and rate) is quite
obvious.

3. Multiphase systems

Multiphase systems combining the properties of two structurally different materials are highly rele-
vant for a number of different technical applications of polymers. Especially for PP, these systems allow
a widening of the property range by combining the advantages of the various phases [114,115]. For
example, most high-impact materials are heterophasic, normally creating problems for optical properties
like transparency and even surface gloss [116]. Multiphase structures can be created in various ways,
both in the polymerisation step and during post-polymerisation modi®cation as will be shown below.
What makes these systems even more interesting from the rheologist's point of view is the interaction
between morphology and rheology: The component rheology determines the developed morphology,
which in turn affects the system rheology and also Ð technically most relevant Ð the mechanical and
optical end-use properties. The subsequent investigation must also account for the sensitivity of such
structures to various probing techniques, especially to deformation.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 921

Fig. 21. Composition dependence of the melt viscosity h of PP/EPDM (PPB) blends as determined in a capillary rheometre at
2008C and a ratio of length/diameter of 20; the wall shear stress t W is: (A) 14 £ 10 4, (B) 18 £ 10 4, (C) 20 £ 10 4 and (D)
22 £ 10 4 Nm 22 (from Ref. [111], reprinted with permission of Elsevier Scienti®c, 2000).

3.1. Inhomogeneous products

The special advantages of polyole®ns with broad or bimodal MMD values have been mentioned
before. Such materials are desirable both for their better processabilities [57] and for their end-use
properties [117] in different areas. As was outlined in Section 2.1, bimodal MMDs can lead to strain-
hardening behaviour and offer mechanical advantages due to the presence of very high molar mass
fractions.
One problem with these products is, however, the potential danger of inhomogeneities within the
material even in case of an identical chemical composition of both high and low molar mass fractions
(e.g. homopolymer). Despite the fact that in the polymerisation step, both fractions are created on the
same catalyst particles, only in serial reactors with different hydrogen contents has, it been found that not
even the polymer powder shows a homogeneous distribution of molar masses over the particles. This
may be due to the multi-site nature of the still mostly used ZN-type catalysts and can possibly be
overcome in case of single-site polymerisations. In any case, the subsequent pelletisation in a twin-
screw extruder is not always suf®cient to render the material homogeneous after production due to the
massive viscosity differences of the fractions (remember: a scaling law with a power of 3±4 applies
between molar mass and viscosity!).
Such inhomogeneity effects can be clearly seen in the rheological behaviour [8], as Fig. 22 shows for
the case of bimodal PP-homopolymer. While a molten powder is still pretty inhomogeneous and has a
rather low overall viscosity due to a dispersion of the high molar mass domains in a low molar mass melt, the
same material has a higher viscosity after homogenisation in a twin-screw extruder with two sets of kneading
blocks. In this particular case, the dispersion has an additional effect of increasing Tc in the DSC experiment
by nearly 4 K, pointing to the nucleating ability of high molar mass tail mentioned in Section 2.1 [69].
922 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 22. Homogenisation effect on storage modulus G 0 (o ) and complex viscosity h p ( 0 ), both from plate/plate geometry at
2308C, of bimodal PP-homopolymer grade; open symbols refer to compressed powder sample, ®lled symbols to compounded
sample.

Other negative effects of inhomogeneity include processing problems like surface distortion and even
melt fracture [17,118] as well as a signi®cant deterioration of the mechanical and optical performance of
these products [119,120]. An important practical aspect is the fact that inhomogeneity can be seen as one
of the main reasons for the appearance of so-called `gels' in both PE- and PP-®lms [121,122]. From the
original de®nition, these are formed by high molar mass or even crosslinked parts (from Latin `gelare').
Generally, however, a differentiation of practically appearing gels by their underlying origins is neces-
sary. Inhomogeneities called `gels' in practice can result from the polymer itself, from decomposition
products and from non-polymeric impurities. These can for example be additive agglomerates or dirt
particles from the production environment (like cellulose ®bres from bags etc.).
Another example for the efforts necessary to achieve homogeneity in a polymer with wide MMD is
given in Fig. 23. Here the dissolution of high molar mass domains (polymer-based gels) in case of a
bimodal HDPE is shown during homogenisation in a Brabender-type two-paddle kneader, where
samples were taken out at different time intervals to check for homogeneities with light microscopy.
It can be clearly seen that only after going through a maximum in energy uptake is homogeneity
achieved. Taking these results for reactor-based systems (where high and low molar masses are already
distributed in the primary particles), it can be easily understood how complicated it is to achieve
extruder-based mixtures with signi®cant viscosity differences between the respective components!

3.2. Filled and reinforced polyole®ns

The main targets of ®lling and reinforcing polyole®ns are twofold: improving the mechanical proper-
ties, mainly stiffness (modulus) and heat de¯ection temperature, and Ð in case of mineral ®llers and
inexpensive organic ®llers mostly Ð also reducing the overall cost of the material [123]. In recent years,
the second target has so lost importance due to the pressure towards weight reduction, especially in the
case of technical applications.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 923

Fig. 23. Homogenisation of a bimodal HDPE in a twin-blade kneader (Brabender Plasticorder); torque curve over time and
optical micrographs (magni®cation 100 £ ) of samples taken at different mixing times.

There is a wide variety of ®llers and reinforcing ®bres presently used for the modi®cation of PP and
(less frequently) PE, which can be categorised as follows [124,125]:

² Mineral ®llers Ð talc, calcium carbonate, kaolin, wollastonite, mica, barium sulphate, etc.
² Natural organic ®llers Ð ®bres of wood, hemp, jute, ¯ax, etc.
² Synthetic organic ®llers Ð ®bres of cellulose and fully synthetic polymers like polyester or aramide,
carbon ®bres
² Glass and other mineral-based ®bres
² Fillers with special purposes Ð iron oxide (magnetism), collodial silver (antibacterial), etc.

In a broad sense, also carbon black and other particulate pigments can be included into this group of
substances, although they are hardly ever present in the material in quantities suf®cient to change the
rheological behaviour of the material.
Relevant parameters of ®llers/®bres are particle size and size distribution, the shape factor (length/
diameter Ð ratio), matrix adhesion, stability against or tendency towards agglomeration, hydrophilic or
hydrophobic natures, the surface polarity and ®nally the purity (mainly with respect to the absence of
metal ions catalysing polymer degradation like iron or copper) of the ®ller. Furthermore, the effects of
additives and processing aids [126,127] have to be considered which can contribute strongly to the
dispersion quality (examples of substances used frequently for this purpose are glycerin-monostearate
and calcium stearate, both amphiphilic in nature).
Another set of relevant parameters comes from mixing and compounding operations. The importance
of a proper selection of mixing equipment and components is given for example in a review by Todd
[128], who stresses the importance of adapting the applied compounding system to the respective task.
The fact that the degree of dispersion additionally affects the rheological properties of a system [129]
allows us to use this property change for an assessment of product quality without the necessity for
924 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

mechanical testing (as will be outlined further below). However, it also makes the system behaviour less
predictable. Compounds are similar to blends discussed in the following sections.
The rheological consequences of ®ller addition are most simple for the Newtonian case, where the
viscosity increase can easily be formulated according to Einstein:
h ˆ hs …1 1 2; 5f† …24†
for the dilute case, where h s is the viscosity of the suspending liquid and f is the volume content of
solids. It is obvious that in this equation neither shape nor size of the dispersed particles are re¯ected.
Any expansion for higher ®ller concentrations will need to consider this. An example is the Krieger±
Dougherty equation for the case of spherical particles which reads
h ˆ hs …1 1 f=fm †2‰hŠfm …25†
with f m being the maximum packing fraction and [h ] is the intrinsic viscosity.
In case of viscoelastic systems, the situation is much more complex. The addition of ®llers here leads
to a change in the relaxation time spectrum with an addition of long relaxation times resulting from
particle-matrix- and particle±particle interactions. In a dynamic-mechanical test this is re¯ected in a
plateau development of the storage modulus at low frequencies, similar to network development in
crosslinking. The extent of this change is affected by all of the above mentioned factors [130±133]. It has
also been described as the development of a yield stress, for which case the stress in shear ¯ow can be
written as
Y
s12 ˆ hg_ ˆ q 1 f …f†G0 teff g_ …26†
1 1 …4=3†t2eff g_ 2

with
t0
teff ˆ …27†
1 1 at0 g_
where Y is the yield stress level, G 0 the modulus and t 0 is the characteristic relaxation time of the un®lled
system and f(f ) a function describing the modulus change by ®lling the polymer with non-interacting
spheres. What makes this formulation somewhat dif®cult is the necessity to determine the value of Y in
an extrapolation of s 12 towards a zero shear rate.
A similar effect will be discussed in Sections 3.4 and 3.5 for the case of polymer blends. This change
of the relaxation behaviour is also re¯ected in elasticity-dominated effects like die swell and melt
fracture [130]. A modi®cation of the ®ller surface will additionally affect the variation [133].
As already mentioned, a correlation of the rheology of ®lled systems to structure and mechanics is
possible to some extent, because of identical in¯uence factors on both properties. This quality control
chance has been tested [134]; the determination of discrete relaxation spectra from the dynamic moduli
of base polymer and compounds (see Fig. 24) allows for the identi®cation of a secondary relaxation
maximum resulting from the ®ller interaction. In Table 3, the results for one compound (which is rather
critical regarding producibility because of the very ®ne talc ®ller used) are given. The values for
t igi(max) were obtained using the BSW-model [45] and discrete spectra with four relaxation times
per decade. Similar correlations can be obtained for a variation of the particle size of the ®ller, where
impact strength and added viscoelasticity also correlate positively.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 925

Fig. 24. Effect of dispersion quality on the rheology of a 30 wt% PP/CaCO3 Ð compound produced on three different
compounding machines (G 0 (v ) from plate/plate measurement at 2308C) compared to base polymer; better dispersion leads
to elasticity at long relaxation times and also improves mechanical properties.

What remains is a problem of measurement, as the materials exhibit solid-like behaviour at low strain
rates (yield stress effect) due to the long relaxation times of the ®ller contribution. This is combined with
a pronounced non-linearity starting at very low stress/strain levels already and a strong response to
sample preparation and pre-deformation effects [135±137]. These effects make it very complicated to
compare results from different laboratories and measuring geometries, as was shown in a round-robin
test [137], in which the reproducibility for G 0 ,G 00 (v )-measurements of un®lled PP was found to be
around ^10%, while for a PP/talc-compound similar to the material discussed in Table 3 a level of
^40% was reached, with special problems at low frequencies, where pre-deformation effectively
destroys the `network effect'. A use of the Cox/Merz-relation as well as the combination of dynamical
and steady-shear results (e.g. capillary data) is virtually excluded.
In practice of production and conversion of such products, also the `side effects' of ®lling in conver-
sion and application need to be considered. These include adsorption of additives like antioxidants [138],
promotion of degradation through a catalytically active ®ller surface, degradation-related die lip bulidup
[20] and promoted melt fracture. Also, defects in molded articles or ®lms are frequently related to

Table 3
Mechanical and rheological parameters of a PP-based talc compound (Daplen KSR 4525 1 30 wt% talc Naintsch A3) re¯ecting
the degree of dispersion (data from Ref. [134])

Extruder Flex. modulus (MPa) Charpy notched 1238C (kJ m 22) git i (max) (Pa s)

ZE 25 `short' 2973 8.8 3.5


ZE 25 `long' 2564 9.8 15
926 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

agglomeration phenomena. The effects of water adsorption promoting the agglomeration tendency of
®llers need to be considered here.
Finally, some special cases should be discussed. Additives and processing aids [121] are a special type
of `®ller' with very small volume fractions in case that they remain solid in the molten polymer. If this is
not the case Ð as for slip agents like erucamide or oleamide, but also for calcium stearate Ð they can
act quite differently on the system behaviour. One of the possible problems appearing in the rheological
testing of polymers containing such `internal lubricants' is a wall slip effect in the measuring instrument.
Pigments and carbon black can have a rheological effect already at very low concentrations due to their
very small particle sizes and Ð especially in case of carbon black Ð good interactions with the
polymer.
Other special cases include network-forming nucleating agents of the sorbitol-derivative type and
nanosised particles, which allow for special perfomances of compounds with respect to mechanics
[139,140] but also barrier properties. Only limited information is available so far on the behaviour of
the resulting `nanocomposites' in rheological tests, but a certain similarity to very ®ne ®llers can be
assumed.

3.3. Crystallizing polyole®ns

As polyole®ns are semicrystalline materials, not only the degree of crystallinity but also the form and
distribution of crystalline structures in the ®nal material have a decisive effect on end-use properties.
This relates to both mechanical and optical performances, on also to long-term stabilities and aging
behaviours.
During the conversion step, rheological properties and the crystallisation process strongly interact
with each other. The complexity of this interaction and the extent to which morphologies are determined
by the rheology of the converted material are de®ned by the type of conversion process.
In their rheological behaviour, crystallizing polymers can be seen as special case of ®lled systems, i.e.
with crystals behaving like a ®ne dispersion of solid particles [141] or as crosslinking systems with the
forming crystals acting as crosslinks between the molecules [142]. Using the ®rst of these assumptions,
one arrives at a formulation like
 
f 22
G 00R …t† ˆ G 00 …t†=G 00 …t ˆ 0† ˆ 1 2 …28†
fm
with G 00 (t) giving the time evolution of the modulus in a crystallisation experiment, f the crystallised
volume and f m Ð as above for ®lled systems Ð the maximum crystalline volume at impingement of
the spherulites. For the remaining crystallisable fraction v(f ), Carrot [141] arrives at
x
…G 0R †21=a ˆ …G 00R †21=2 ˆ v…f† ˆ 1 2 t …29†
x1
where x de®nes the degree of crystallinity. As the long relaxation times are affected most by crystal-
lisation, the measurement is most sensitive at low frequencies.
This opens an interesting possibility to monitor crystallisation processes with rheological methods. An
inherent problem in such tests is the in¯uence of the strain-induced creation of nuclei, effectively the ®rst
step of shear-induced crystallisation. If the quiescent crystallisation should be studied in an undisturbed
form [142, 143], only very low stresses and reversible deformations like in a dynamic test can be applied.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 927

Fig. 25. Development of storage modulus G 0 and tangent of loss angle tan(d ) for iPP-homopolymer with MW ˆ 500 kg mol21 ;
MW =MN ˆ 5:0 during a quench to 1388C after melting at 2608C and subsequent shearing during the indicated times at
g 0 ˆ 5 s21 (from Ref. [67]; reprinted with permission of Steinkopff Verlag, 2000).

However, a combination of unidirectional and dynamic shear tests also allows for determination of the
sensitivity of a material to deformation- or stress-induced crystallisation. Such investigations have been
published for iPP [67,144±146] and poly-1-butene [147]. An example for iPP is given in Fig. 25, it
clearly shows the acceleration of the crystallisation process through a pre-shearing step.
Especially for the case of iPP, it can be seen from the quoted literature that the MMD affects both
rheology and crystallisation behaviour. This is also re¯ected in processing behaviour and morphology
formation [66]. The fact that polymers with a higher molar masses and polydispersities show signi®-
cantly more oriented structures (like shear-induced skin layers in injection moulding) is a direct conse-
quence of this interaction [65]. It is further enhanced by the `gelation effect' of forming crystallites (see
above), which makes the process self-accelerating.

3.4. Elastomer blends

This section covers a wide area of technically applied systems. Generally, elastomer blends are not
limited to polyole®n-based systems, and the principle of impact modi®ed materials is applied also in
chemically different material classes like polystyrene on polyamide. The combination of stiff matrix
materials with ¯exible, particular inclusions being able to absorb stresses in case of deformation and
preventing crack propagation and failure is a general design principle for many synthetic materials.
This area will be used here to outline the principles of polymer blend theory, which is based mainly on
the work of Utracki [148,149]. The phase structure is mainly a function of the relationship between the
viscosities of the disperse and matrix phases

l ˆ hDisp: =hMatrix …30†


928 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 26. Effect of dispersion (B) and agglomeration (C) processes on the particle size distribution of a 2-phase polymer system.

and the capillary- or Weber-number of the mixing process


We ˆ …g 0 hMatrix R†=a …31†
where R is the (original) radius of the disperse phase particles and a is the interfacial tension between the
phases. In the mixing process, additional particle breakup only takes place in case of We . Wecrit, with
Wecrit being determined by l and normally reaching a realistic value only for 0.1 , l , 1. This concept
has several weaknesses. Firstly, it only considers dispersion (breakup) processes and no agglomerations,
which play an important role in structure development as well. The ®nal particle size distribution as
outlined in Fig. 26 is always rather wide, as the stabilty limit for the particle diameter is different for both
processes. This equilibrium has been investigated in the group of Moldenaers [150±152] using rheolo-
gical techniques for the characterisation of polymer blends, both model systems and technically relevant
ones. An important fact for this development is the stability of very small particles which resist agglom-
eration due to their high surface tensions.
For the rheological investigation of such blend systems the same caution as mentioned before for ®lled
polymers is necessary. As for ®llers, the shape and size of the dispersed particles both in¯uence the
system behaviour. Unlike ®ller particles these can be massively changed in a ¯ow ®eld. Again, only an
application of low and dynamic stresses and strains makes the results representative for the equilibrium
structure of the investigated material.

3.4.1. Extruder blends


Different types of elastomers with or without crosslinking are in practical use for the modi®cation of
polyole®ns. Roughly, these can be separated into

² Ole®n-based types like amorphous ethylene-propylene copolymers (EPR), ethylenene-propylene-


diene rubbers (EPDM) and single-site catalyst based plastomers (MC-LLDPE).
² Non-ole®n based types like styrene-ethylene-butadiene rubbers (SEBS), natural rubber (NR), buta-
diene rubber (BR) and ethylene-vinylacetate-copolymers (EVA).

The second category also adds polarity to the blends is normally used only if special properties, e.g.
polarity, are required, as the higher cost of these components prevents a more general application.
Again, the design parameters include not only the components and mass ratios selected for the blend,
but also the mixing equipment used. Mostly used are twin screw extruders with special mixing sections
(e.g. kneading blocks, distribution disks), but also other types are in use. For EPR- or EPDM-types
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 929

Fig. 27. Capillary rheology of PP/EPDM-blends (2008C) based on a PP-homopolymer (B) and EPDM CO038 (X); blends with
20 wt% (A), 40 wt% (L), 60 wt% (K) and 80 wt%(W) of EPDM (from Ref. [153]; reprinted with permission of Elsevier
Scienti®c, 2000).

supplied in bales instead of granules (very low crystallinity grades), batch mixers similar to a Brabender
twin-blade kneader are in use (possible in combination with a granulation extruder like in a Farell-type).
Also, the co-kneader geometry of Buss is frequently used, which has shown advantages for feeding
liquid components like peroxides for partially crosslinked systems.
An important difference between crosslinked and non-crosslinked systems is the different behaviour
of the particles in ¯ow. While non-crosslinked elastomers will deform and eventually break up in a ¯ow
®eld [153], the deformation will be followed by a recoil. Thus, systems with crosslinked elastomers like
thermoplastic vulcanisates (TPVs) are much less sensitive to processing effects [111±113,154]. These
phase effects are re¯ected in the rheological behaviour only to a limited extent, especially if steady-state
shear properties like in a capillary experiment (for example see Fig. 27) are investigated.
Mostly, the morphology development is determined by the phase viscosities. An example for PP/
EPDM-mixtures is given in Table 4 and Fig. 28, where the viscosity of the matrix PP was varied
systematically, resulting in different particle sizes and impact strengths for the resulting systems. The
formation of an equlibrium particle size distribution in such systems by breakup and coalescence
processes, as outlined before, is described in the literature [150,151].
An additional effect of phase compatibility, expressed in the interfacial tension a between the matrix
and elastomer phase, has been shown for PP/EPR-systems [155,156] and other elastomers like MC-
based plastomers [157,158]. An example for this effect is the improvement of both phase structure and
mechanical properties by changing the matrix of a PP/EPR-blend from a homopolymer to an EP-random
copolymer. Some results of a recent development study [156] are shown in Table 5 and Figs. 29 and 30.
This principle of improved phase compatibility, the importance of which will become even more evident
in case of blends with chemically different polymers, has been applied in the development of copolymers
with an improved transparency/toughness ratio [116]. The closer chemical and mechanical similarity
930 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Table 4
Effect of viscosity ratio between matrix and elastomer phase on morphology and mechanics of PP/EPDM-blends (base
polymers: Daplen PP; elastomer: Dutral CO 038 Ð viscosity approx. identical to DS 10)

Material Matrix Matrix MFR (g/10 min) Flex. test (MPa) IZOD notched (kJ m 22) Particles (mm)

2308C/2,16 kg Modulus 1238C 08C 2208C Diam.

4512/01 BE 50 0.3 1110 58.2 5.71 3.88 0.2±0.6


4512/02 DS 10 2.4 1145 6.43 3.34 3.21 ±
4512/03 KS 10 8 1070 4.18 2.42 2.87 1.0±3.0
4512/04 MT 58 13 1140 3.66 2.55 2.60 ±
4512/05 RT 58 25 1155 2.90 2.00 2.06 10±50

between the phases leads to an additional improvement of the stress whitening behaviour due to a change
of the micro-deformation processes in the material. The combination of smaller elastomer particles and a
reduction of the modulus difference between matrix and dispersed phase results in a transition from
crazing to shear yielding, eliminating the formation of microvoids at already low deformations.
For a detailed analysis of the rheological effects, the interfacial contributions in blend rheology
have to be discussed. A blend system without signi®cant interactions between the phases should be
characterised by a logarithmic behaviour like
X
fi log…Gpj …v††
p
Gmix …v† ˆ 10 i …32†

for the complex modulus G p of the blend, where the Gip are the moduli and f i are the volume fractions of
the components. This formulation approximates the rheology of PP/EPR-systems quite closely, as can be
seen in Fig. 31. If interface tension or particle/particle-interactions contribute to the system behaviour,
both positive and negative deviations from Eq. (32) can be found. The addition of long relaxation times
and elasticity to the system [159,160] will be discussed in latter. For pure PO-systems, also viscosity
depression in mixtures has been reported [161]. In any case, knowing the phase morphology is very

Fig. 28. Effect of viscosity ratio between matrix and elastomer phase on the impact strength above and below the glass transition
of the matrix; for details of the material composition see Table 3.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 931

Table 5
Morphological and mechanical characteristics of model compounds and a reactor product comparable to mat. 8825/01; effect of
increased compatibility in case of copolymer matrix (data from Ref. [157])

Material Matrix Avg. particle size (mm) Flex. modulus Charpy U-notch
1238C (MPa) 1238C (kJ m 22)

8825/04 KS 101 2.5 1085 8.3


8825/01 KFC 2004 1.3 885 11.8
8825/02 KFC 2006 1.0 737 14.3
8825/03 KFC 2008 0.8 618 15.1
6473/02 Reactor product 0.25 604 20.1

important if the rheological data should be used for the calculation of system parameters. Here, it is
important to distinguish between particle and pro®le size distribution [162].
Rheological studies of PO-blends also allow correlations to the processing behaviour. The hetero-
phasic structure can result in a signi®cant elasticity increase, surface defects and melt fracture [163].
Also, the correlation to ®nal properties is possible within limits. Both particle size distribution and EPR
fraction are re¯ected in the rheology of the systems [164,165]. As shown in Figs. 32 and 33, which come
from a dilution series of a reactor-blend as described before [116] made for a systematic variation of the
interparticle distance at constant particle size [165], a higher elastomer content is re¯ected both in
mechanical and rheological behaviour.
In industrial practice, frequently a combination of EPR and HDPE is used for the modi®cation of PP,
resulting in three-phase systems [166±168]. In contrast the present trend towards application of pure
MC-plastomers Ð mainly because of their availability in a wide range of MFR (MMD) and density
(comonomer content) Ð results in two-phase systems [169,170].

Fig. 29. Dynamic viscosity (calculated from plate/plate-test at 2308C) and viscosity ratio for the system of Table 5 and Fig. 30;
PP matrix (A) Ð Daplen KFC 2006 Ð and EPR (W) Ð Exxelor VM42E.
932 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 30. Morphology of blends from Table 5; KS 101/VM42 (left) and KFC 2008/VM42 (right).

3.4.2. Reactor blends


Similar application properties can be achieved by the production of heterophasic EP-copolymers,
which have a signi®cant economical advantage over extruder-based systems. The polymerisation
sequence of various fractions in a reactor cascade and the volume ratio between the same allows for
a wide range of possible product characteristics. The accessible product range can be further expanded
nowadays with new MC-catalysts through incorporation of new monomers and improved distribution of
the comonomers [6,171].
As for the extruder-based blends, the phase structure dominates the behaviour of heterophasic copo-
lymers. These are basically three-phase systems, the structure development from the reactor powder is
shown in Fig. 34. The system rheology is also very similar to extruder blends [155,172]. Details of this

Fig. 31. Rheology of PP/EPR-blend from Table 5 and Fig. 29 (KFC 2006/VM42E) compared to sum of components (loga-
rithmic mixing rule) Ð plate/plate 2308C.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 933

Fig. 32. Dynamic viscosity of pure RAHECO SVA-279 (dashed line), mixture with 30 wt% matrix polymer (A) and pure matrix
polymer KFC 2008 (solid line); from dilution series in Fig. 33.

development are important because of the structure effects on the toughness, as a consequence of
microdeformation processes ocurring at the interface matrix-elastomer and inside the elastomer particles
[173].
Also, these materials open additional modi®cation possibilities through a combination with other
components in extrusion mixing. Different elastomers, but also HDPE and ®llers can be added, resulting
in very complex compounds. This kind of property design [115] is of great importance for technically
challenging applications. An interesting trend here is the partial substitution of mineral-®lled compounds
with high crystallinity PP matrices, resulting in various advantages like weight reduction, better
organoleptic properties and improved scratch resistances.
Important for PP/EPR-systems, both from extruder and reactor, are degradation effects in the

Fig. 33. Elastomer concentration effect on stiffness and impact strength at identical particle size distrbution; dilution series
based on experimental RAHECO-grade SVA-279 with KFC 2208 as matrix component.
934 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Fig. 34. Phase structure development in a heterophasic EP-copolymer; starting from the reactor particle in a melting step (A) a
primary morphology develops, which can change in a high strain environment (B) by further breakup and dispersion, in a low
strain environment (C) by agglomeration (see also Fig. 23).

CR-process or even in normal compounding or conversion steps. The different effect of radicals or
peroxide on matrix and elastomer (discussed before in Section 2.1) leads to a change in phase structure
and properties [174]. The composition and structure of the original system allows a partial compensation
of these negative effects, as can be seen for three different degradation series from Table 6. This is
closely connected to the rheology and morphology of the materials as shown in Figs. 35 and 36; an
extreme consequence can be the formation of crosslinked gels.
The application areas are strongly related to the mechanical pro®le of these materials, with low-
temperature impact strength as the most important property. Also other relevant application properties
like surface replication etc. are related to phase structure and component properties, also of the exter-
nally added components.

Table 6
MMD development in the degradation of the three polymer types for total polymer and xylene solubles (XS) content (HOMO
Ð homopolymer, HECO Ð heterophasic copolymer with 12 mol% ethylene, RAHECO Ð random-heterophasic copolymer
with 25 mol% ethylene; data from Ref. [174])

Material MFR (g/10 min) XS (wt%) GPC total (kg mol 21) GPC(XS) (kg mol 21)

2308C/2,16 kg MW MW/MN MW

B-HOMO 0.5 ± 1061 6.4 ±


BHD2 5 ± 322 3.4 ±
BHD3 50 ± 196 3.5 ±
B-HECO 0.5 12.0 1069 5.6 532
BCD2 5 13.2 322 3.2 267
BCD3 50 13.0 152 3.0 158
B-RAHECO 0.5 25.0 667 4.9 534
BRD2 5 27.7 266 3.0 263
BRD3 50 26.1 124 2.4 157
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 935

Fig. 35. Evolution of the rheological behaviour in the HOMO- (left) and the RAHECO- (right) degradation series of Table 5
(solid lines Ð storage modulus G 0 , dashed lines Ð complex viscosity h p).

3.5. Blends with other polymers

Different principles apply for blends between chemically very different polymers. Here, an effective
compatibilisation between the main phases by a third component is necessary to optimize the morphol-
ogy, and to ensure processing stability and appropriate mechanics [148]. Essentially, a blend compati-
bilizer has two functions: reducing the interfacial tension in the melt phase for achieving an optimum
phase structure, and improving the phase interaction in the solid phase.
Before going to practical examples, the principles of morphology development and rheological
properties of PO-blends with non-ole®nic polymers will be outlined brie¯y. In principle, the behaviour
is similar to Sections 3.1 and 3.2, although a much stronger interface effect and non-linearity is found,
resulting in a high relevance of predeformation and non-linearity effects as in Section 3.2 [176]. For
describing the rheology, an interfacial term GpIF as in
X
fi log…Gpj …v††
p
Gmix …v† ˆ 10 i 1 GpIF …33†

Fig. 36. Effect of degradation on morphology in the RAHECO-series of Table 5 and Fig. 31 (left Ð MFR 0.5, right Ð MFR
50); TEM-pictures after RuO4-staining (magni®cation approx. 2.500 £ ).
936 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

has to be added. Different formulations have been proposed for analysing this component to gain
additional system information [159,160]. In practice, the method of Gramespacher and Meissner
[160] has proven quite successful, in which the characteristic relaxation time t IF for GpIF is determined
as described before using a discrete relaxation time spectrum. If the average particle radius R of the
system is known (e.g. from SEM- or TEM-investigations as described in Ref. [162]), the interfacial
tension a can be calculated from

hM R …19k 1 16†=…2k 1 3†
tˆ with k ˆ hD =hM …34†
a 40…k 1 1†

where h D and h M are the viscosities of the disperse and the matrix phase, respectively. Both this method
and the dispersion model developed by Palierne [159,177] are however limited to systems with a rather
low contents of disperse phase, as otherwise co-continuous structures are developed, for which no clear
particle dimension can be determined.
Practical examples, which demonstrate the importance of component selection, application of compa-
tibilizers and blend production will be discussed for two pairs of polymers:

² Polypropylene/Polyamide-6 (PP/PA-6). This system was very well investigated in the late 80s and
early 90s by various companies, although little of the developments made were ultimately commer-
cialised. Literature data are available [175,176,178], discussing composition effects on morphology,
rheology and mechanics of these blends. Different types of compatibilizers were used in these studies,
mostly polyole®ns Ð PP or EPR Ð grafted with maleic anhydride (MAH) to ensure compatibility
with the PA-6 phase. Another class of promising polymers for this purpose were styrene-elastomers
like styrene-ethylene-butadiene-copolymers (SEBS) grafted with MAH, which showed excellent
behaviour regarding blend mechanics due to their elastomeric nature.
² An example of the rheology and morphology of such a system based on PP-homopolymer, PA-6 and
SEBS-g-MAH is given in Figs. 37 and 38. The comparison between the calculation result according
to Eq. (32) and the measured data for the storage modulus in Fig. 37 clearly demonstrates the
mentioned interfacial contribution. A rather ®ne morphology is achieved with the applied amount
of 5 wt% of compatibilizer; a close look at Fig. 38 shows that this is reached with a more or less
complete coverage of the dispersed PA-6 particles with the SEBS-g-MAH. The consequences of
selecting type and quantity of compatibilizer on the mechanics of PP/PA blends becomes clear from a
look at Table 7, where different blends with identical main components and PA-contents are
compared.
² Polyethylene/Polystyrene (PE/PS). For this system, some more recent examples have been published
[179,180]. Despite its limited thermomechanical stability …TG ˆ 90±1008C†; PS is an interesting
blend partner for PO's from mechanical reasons as well as paintability and printability. Styrene-
elastomers like SEBS can be used as compatibilizers, but also reactive blending using styrene/
peroxide Ð combinations have proven to be viable. High impact strengths for these systems can
only be reached with the incorporation of an additional elastomeric phase, prefrentially at the
interface.

In any case, the development of such blends must be targeted at properties not achievable in pure
PO-based systems to justify the certain additional effort and cost. For this, also a combination with
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 937

Fig. 37. Rheology of a PP/PA-blend (75 wt% PP Daplen BE 50, 20 wt% PA Durethan B30S, 5 wt% SEBS-g-MAH); storage
modulus of components, calculated curve with moxing rule and measured curve (2308C, plate/plate after drying).

elastomers is of interest, possibly already as compatibilisers, and ®llers for creation of network-like
structures are possible.
Despite the original expectations, the application of PO-blends with non-ole®nic polymers has
remained limited to applications in which properties like polarity, high HDT, HF-weldability and barrier
properties are important. Also reactive blends dominating the market in which the second phase and the
compatibilizer are produced in one step, leading to improved process and product economics.

Fig. 38. Morphology of PP/PA-blend as in Fig. 37.


938 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Table 7
Mechanical properties of PP/PA-blends as in¯uenced by compatibiliser type and quantity (all properties determined on
injection-moulded plates 150 £ 80 £ 2 mm 2 parallel to injection direction; FWI Ð falling weight impact, Charpy DV Ð
double V-notch)

M 5969 M 5976 M 5972 M 5984

PP type Daplen BE 50 Daplen BE 50 Daplen BE 50 Daplen BE 50


PA type ± Durethan B30S Durethan B30S Durethan B30S Durethan B30S
PA quantity wt% 20 20 20 20
Comp. type ± None Kraton G 1901X Kraton G 1901X Admer GR2
Comp. quantity wt% 0 1 5 5
MFR g/10 min 2.03 0.96 0.24 0.33
Flex. modulus MPa 1420 1540 1560 1480
FWI (Wtot) J 1.5 2.3 43.9 4.9
Charpy DV kJ m 22 7.6 11.0 45.2 42.8

4. Actual and future trends

A basic requirement for following future trends in the polymer industry is to understand the principles
of market growth as shown for the case of PP in Fig. 39. More than 75% of the volume growth of
polyole®ns results from intermaterial substitution, still more than 50% from inter-polymer substitution.
One of the main developments during the 90s was the replacement of `technopolymers' (or `engineering
thermoplastics', ETPs) like polyamides or ABS with polyole®ns, especially by PP. This cost-driven
process was partly facilitated by constructional improvements within the concerned parts, but mostly by
signi®cant expansions of the property range of these polyole®n materials.
There is more to come; several upcoming developments will ensure a continuation of this trend over
the next decade:

² Metallocene catalysts, which have already had a major impact for PE [181] and elastomers will gain
in importance for PP [18]. The key advantages of this catalyst family like narrow MMD, possibility to
incorporate new comonomers and to achieve a better randomness, generally better control of the
polymer structure Ð will allow for an even wider property range. Although the speed of market
introduction has so far been lower than expected (partially the process was retarded by legal
problems), this development will remain among the dominating ones.
² New polymerisation technologies giving access to a wider range of products [70] are constantly
gaining ground. The possibilities of polymerisation under supercritical conditions have so far not
been fully exploited for the market. A problem for widening the application range may, however,
arise from the increasing plant size, which makes it virtually impossible to produce small-volume
grades for market niches. On-line production control by rheological techniques will gain in
importance.
² Ultrahigh MFR materials, which have good mechanical properties, are a development dictated by the
market. New and faster processing technologies as well as wall thickness reduction (downgauging)
for economical and ecological reasons are the underlying reasons. Here, present rheological tech-
niques are challenged by the low viscosity of these materials.
² Controlled LCB systems, produced via polymerisation and/or reactive compounding, should allow
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 939

Fig. 39. Market growth of PP in Western Europe 1993±1998 segmented by growth motivators (economical development, new
applications and material substitution).

the access of PP to areas presently occupied by glassy polymers. Beyond extrusion foaming, there is
also ISBM as a potential growth area. As outlined in Section 2.2.2, rheological means of material
characterisation are most important here.
² New ®ller types, especially organic ®bres, which presently show an above average market growth,
may help property development in various ways. For one, the ecological factor of sustainability must
be considered, and weight reduction will remain an important argument. Also, energetic recycling is
facilitated.
² Nanocomposites, which are already now a very interesting material family, have not yet demonstrated
their full range of possibilities. For a more widespread market penetration, however, signi®cantly
cheaper ways of production will have to be found. The rheological properties of these materials are
challenging as has been shown already [139,140].
² Selective crosslinking in multiphase systems, which already helps in optimizing property pro®les,
will gain in importance as one of the possible modi®cation technologies allowing for the production
of a multitude of tailor-made grades for market niches that are based on one reactor grade of polymer.

A much used, but less often considered tool in speeding up market-oriented product development is
the so-called chain of knowledge, which should ideally stretch from the catalyst to the long-term
behaviour of the ®nal part, and possibly even include the recycling possibilities. One important pre-
requisite for a growing understanding of interactions between different steps of the production chain and
also onward into conversion and application is the necessity to `talk one language' throughout polymer
development. Some of the requirements seem trivial Ð like the consistent use of SI units in reporting
and publishing Ð some other are less so. Interdisciplinary efforts for creating new, value-added materi-
als will have the biggest chances for success in years to come and should be promoted especially in
academia and education.
940 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

Acknowledgements

I would like to thank several people, who have contributed to the creation of this Review: In the
startup phase, Prof. Otto Vogl, New York, for encouraging me to write it as well as Prof. Manfred
Wagner, Berlin, and Prof. Alois Schausberger, Linz, for helping me to the right selection of literature.
Then my fellow rheologists in this company for supplying data: Svein Eggen from Rùnningen, Anneli
Malmberg from Porvoo, Tonja Schedenig from Schwechat and Bernhard Knogler from Linz. And at last,
all of my colleagues who helped me ®nishing the work with discussions, corrections and hints Ð
especially Norbert Reichelt and Klaus Bernreitner in Linz and Pirjo JaÈaÈskelaÈinen in Porvoo.

References

[1] Macosko CW. Rheology: principles, measurements and applications. New York: VCH Publishers, 1994.
[2] Dealy JM, Wissbrun KF. Melt rheology and its role in plastics processing. London: Chapman & Hall, 1995.
[3] Gartner C, Sierra JD, Avakan R. Proc ANTEC '98, Soc Plastics Engineers, 1998. p. 2012±6.
[4] Yan D, Wang W-J, Zhu S. Polymer 1999;40:1737±44.
[5] Bersted BH, Slee JD, Richter CA. J Appl Polym Sci 1991;26:1001±14.
[6] Moore EP, editor. Polypropylene handbook. Munich: Hanser, 1996.
[7] Aenderkerk J, Michels D. In: Proceedings of Metallocenes '96 Conference, Schotland, DuÈsseldorf, 1996, 4/01-12.
[8] Gahleitner M. J Macromol Sci Pure Appl Chem A 1999;36:1731±41.
[9] Gabriel C, Kaschta J. Rheol Acta 1998;37:358±70.
[10] Bremner T, Rudin A, Cook DG. J Appl Polym Sci 1990;41:1617±27.
[11] Holland D. J Rheol 1994;38:1941±3.
[12] Lomellini P, Ferri D. In: Proceedings of 13th International Conference on Rheology, Cambridge, 2000, 1-118-
120.
[13] Sammler RL, Koopmans RJ, Magnus MA, Bosnyak CP. Proc ANTEC '98, Soc Plastics Engineers, 1998. p. 957±61.
[14] Gotsis AD, Ke Q. Proc ANTEC '99, Soc Plastics Engineers, 1999. p. 1156±60.
[15] Rides M, Allen CRG, Chakravorty S. In: Proceedings of PPS 15 Conference, 's Hertogenbosch/NL 1999, #224.
[16] Bidell W, Hingmann R, Jones P, Langhauser F, Marczinke B, Mueller P, Fischer D. In: Proceedings of Metallocenes '96
Conference, Schotland, DuÈsseldorf, 1996, 4-26-34.
[17] Abbas K, Heino EL, Kjeldsen K, Asrud P. In: Proceedings of PPS Europe/Africa Regional Meeting, Polymer Processing
Society, Gothenburg, 1997, PL1.
[18] Friedrich C, Eckstein A, Stricker F, Muehlhaupt R. In: Scheirs J, Kaminsky W, editors. Metallocene-based polyole®ns.
vol. 2. p. 399±416.
[19] Fleissner M. Int Polym Proc 1988;2:229±33.
[20] Gander JD, Giacomin AJ. Polym Engng Sci 1997;37:1113±26.
[21] Andreassen E, Bùrve KL, Rommetveit K, Redford K. Proc ANTEC '99, Soc Plastics Engineers, 1999. p. 2104±8.
[22] Phillips R, Herbert G, News J, Wolkowicz M. Polym Engng Sci 1994;34:1731±43.
[23] Gahleitner M, Fiebig J, Paulik C, Wolfschwenger J, Zuidema H, Dreiling G, Proc. of PPS 15 Conference, 's Hertogen-
bosch/NL 1999, # 008 (CD-ROM Edition).
[24] Ryan EM, Garcia-Rejon A. Proc ANTEC '94, Soc Plastics Engineers, 1994. p. 1824±30.
[25] Fulchiron R, Verney V, Michel A, Roustant JC. Polym Engng Sci 1995;35:513±7.
[26] Vlachopoulos J, Bellehumeur CT, Kontoupoulou M. In: Proceedings XIIth International Congress on Rheology, Cana-
dian Rheology Group, 1996. p. 693-4.
[27] Kontoupoulou M, Vlachopoulos J. Polym Engng Sci 1999;39:1189±98.
[28] Hamielec AE, Soares JPB. Prog Polym Sci 1996;21:651±706.
[29] Hammerschmid K, Gahleitner M. In: Karger-Kocsis J, editor. Polypropylene Ð an A to Z reference. Dordrecht: Kluwer
Academic Publisher, 1999. p. 95±103.
[30] Tzoganakis C, Vlachopoulos J, Hamielec AE. Polym Engng Sci 1988;28:170±80.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 941

[31] Mendes L, D'Alessandro J, Gomes A, Lachtermacher M. In: Proceedings of ANTEC 2000, Society of Plastics Engineers,
2000, #0967 (CD-ROM Edition).
[32] Moad G. Prog Polym Sci 1999;24:81±142.
[33] Valenza A, Piccarolo S, Spadaro G. Polymer 1999;40:835±41.
[34] Lazar M, Hrckova L, Borsig E, Marcinin A, Reichelt N, RaÈtzsch M. J Appl Polym Sci 2000;78:886±93.
[35] Cramez MC, Oliveira MJ, Crawford RD. Proceedings of ANTEC 2000, Society of Plastics Engineers, 2000, #0099 (CD-
ROM edition).
[36] Zeichner GR, Patel PD. In: Proceedings of Second World Congress of Chemical Engineering, Society Chemical
Engineers Montreal, 1981. p. 333±7
[37] Bernreitner K, Neissl W, Gahleitner M. Polym Testing 1992;11:89±100.
[38] Yoo HJ. Adv Polym Technol 1994;13:201±5.
[39] Wasserman SH. Proc ANTEC '97, 1997. p. 1129±33.
[40] Steeman PAM. Rheol Acta 1998;37:583±92.
[41] Jordens K, Wilkes GL, Janzen J, Rohl®ng DC, Welch MB. Polymer 2000;41:7175±92.
[42] Gahleitner M, Knogler B, Hafner N, Pirgov W. In: Proceedings of 13th International Conference on Rheology,
Cambridge, 2000, 1-121-3.
[43] Minoshima W, White JL, Spruiell JE. Polym Engng Sci 1980;20:1166±76.
[44] Montfort JP, Marin G, Monge P. Macromolecules 1986;19:1979±88.
[45] Baumgaertl M, Winter HH. Rheol Acta 1989;28:511±9.
[46] Mead DW. J Rheol 1994;38:1769±75.
[47] Brabec C, Schausberger A. Rheol Acta 1995;34:397±405.
[48] Chambon F. Proc ANTEC '95, Soc Plastics Engineers, 1995. p. 1157±61.
[49] Carrot C, Revenu P, Guillet J. J Appl Polym Sci 1996;61:1887±97.
[50] Wasserman SH, Graessley WW. Polym Engng Sci 1996;36:852±61.
[51] Anderssen RS, Mead DW, Driscoll IV JJ. J Non-Newt Fluid Mech 1997;68:291±301.
[52] Eggen S, Karstang T. In: Proceedings of the 12th International Congress Rheology, Quebec City, 1996. p. 669±71.
[53] Eggen S, Bryntesen H. In: Proceedings of the 13th International Conference on Rheology, Cambridge, 2000, 1-127-9.
[54] Ringhofer M, Gahleitner M, Sobczak R. Rheol Acta 1996;35:303±7.
[55] Ringhofer M, Brabec C, Sobczak R, Mead D, Driscoll J. Rheol Acta 1997;36:657±66.
[56] Wagner MH, Ehrecke P, Hachmann P, Meissner J. J Rheol 1998;42:621±38.
[57] Wolfsberger A. In: Proceedings of Polypropylene '93, ZuÈrich, Maack Business Services. 1993, III/1 (1-29).
[58] Eggen S, Sommerfeldt A. Polym Engng Sci 1996;36:336±46.
[59] Khare A, Qin C, Ling MTK, Woo L. Proc ANTEC '98, Soc of Plastics Engineers, 1998. p. 2148±51.
[60] Shroff R, Prasad A, Lee C. J Polym Sci B: Polym Phys 1996;34:2317±33.
[61] Tzoganakis C, Vlachopoulos J, Hamielec AE. Polym Engng Sci 1989;29:390±6.
[62] Bailey PN, Varral DC. In: Proceedings of Metallocenes '96 Conference, Schotland, DuÈsseldorf 1996 1/5 (1±12).
[63] Huang Y-L-, Brown N. J Polym Sci B: Polym Phys 1991;29:129±37.
[64] Brown N, Zhou Z. Macromolecules 1995;28:1807±11.
[65] Gahleitner M, Wolfschwenger J, Bachner C, Bernreitner K, Neissl W. J Appl Polym Sci 1996;61:649±57.
[66] Fujiyama M, Wakino T. J Appl Polym Sci 1991;43:57±81.
[67] Vleeshouwers S, Meijer HEH. Rheol Acta 1996;35:391±9.
[68] Cheng CY, Kuo JWC. Proc ANTEC '97, Soc of Plastics Engineers, 1997. p. 1942±9.
[69] Gahleitner M, Bachner C, Ratajski E, Rohaczek G, Neissl W. J Appl Polym Sci 1999;73:2507±15.
[70] Takakarhu J. In: Proceedings of Polypropylene '99, Maack Business Services, 1999, 1/3 (1±7).
[71] Wadud SEB, Baird DG. Proc ANTEC '99, Soc Plastics Engineers, 1999. p. 1200±4.
[72] Malmberg A, Liimatta J, Lehtinen A, LoÈfgren B. Macromolecules 1999. p. 32:6687±96.
[73] Fraser WA. In: Proceedings of Metallocenes '96, DuÈsseldorf '96, 1996. p. 59±86.
[74] Winter HH, Holly E. Proc ANTEC '85, Soc Plastics Engineers, 1985. p. 613±5.
[75] Micic P, Bhattacharya SN, Field G. Int Polym Proc 1996;11:14±20.
[76] Eckstein A, Friedrich C, Lobbrecht A, Spitz R. Acta Polym 1997;48:41±6.
[77] Eckstein A, Suhm J, Friedrich C, Maier R-D, Sassmanshausen J, Bochmann M, Muehlhaupt R. Macromolecules
1998;31:1335±40.
942 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

[78] McLeish TCB, Larson RG. J Rheol 1998;42:81±109.


[79] Inkson NJ, McLeish TCB, Harlen OG, Groves DJ. J Rheol 1999;43:873±96.
[80] McLeish TCB. In: Proceedings of PPS 15, 's Hertogenbosch 1999, #373 (CD-Rom edition).
[81] Starck P, Lindbergh J. Angew Makromol Chem 1979;75:1±27.
[82] DeMaio VV, Dong D. Proc ANTEC '97, Soc Plastics Engineers, 1997. p. 1512±6.
[83] Rokudai M. J. Appl Polym Sci 1979;23:463±71.
[84] Ram A, Izrailov L. J Appl Polym Sci 1986;31:85±100.
[85] Schedenig T, Schausberger A, Gahleitner M, Pirgov W. Proc Fifth Eur Rheol Conf, Protoroz/SLO 1998:307±8.
[86] Kim YC, Yang KS, Choi C-H. J Appl Polym Sci 1998;70:2187±95.
[87] Deeprasertkul C, Rosenblatt C, Wang S-Q. Macromol Chem Phys 1998;199:2113±8.
[88] Hong Y, Coombs SJ, Cooper-White JJ, Mackay ME, Hawker CJ, MalmstroÈm E, Rehberg N. Polymer 2000;41:7705±13.
[89] Ghijsels A, Massardier CHC, Bradley RM. Int Polym Proc 1997;12:147±54.
[90] Laun HM, Muendstedt H. Rheol Acta 1978;17:415±27.
[91] Laun HM, Schuch H. J Rheol 1989;33:119±75.
[92] Mackay ME, Dajan AM, Wippel H, Janeschitz-Kriegl H, Lipp M. J Rheol 1995;39:1±14.
[93] Jacovic M, Pollock D, Porter R. J Appl Polym Sci 1979;23:517±27.
[94] Dealy JM, Wood-Adams P. In: Proceedings of PPS America Regional Meeting, Polymer Processing Society, Toronto,
1998. p. 3±4
[95] Malmberg A, Gabriel C, Stef¯ T, LoÈfgren B. In: Proceedings of 13th International Conference on Rheology, Cambridge,
2000, 1-174-6.
[96] RaÈtzsch M, Bucka H, Panzer U. In: Karger-Kocsis J, editor. Polypropylene Ð an A to Z reference. Dordrecht: Kluwer
Academic Publisher, 1999. p. 635±42.
[97] Sugimoto M, Tanaka T, Masubuchi Y, Takimoto J-I. J Appl Polym Sci 1999;73:1493±500.
[98] Tzoganakis C. Can J Chem Engng 1994;72:749±54.
[99] Kurzbeck S, Oster F, Muendstedt H, Nguyen TQ, Gensler R. J Rheol 1999;43:359±74.
[100] Park CB, Cheung LK. Polym Engng Sci 1997;37:1±10.
[101] Wolfsberger A, Gahleitner M, Niedersuess P. In: Proceedings of Special Polyole®ns '99, Schotland Business Services.
Houston/TX, 1999. p. 201±14.
[102] DeMaio VV, Dong D, Gupta A. In: Proceedings of ANTEC 2000, Society of Plastics Engineers, 2000, #0236 (CD-ROM
edition).
[103] Wagner MH, Collignon B, Verbeke J. Rheol Acta 1996;35:117±26.
[104] Wagner MH, Bernnat A, Schulze V. J Rheol 1998;42:917±28.
[105] Linster JJ, Meissner J. Polym Bull 1986;16:187±94.
[106] Koopmans RJ. Proc ANTEC '97, Soc Plastics Engineers, 1997. p. 1006±9.
[107] Muendstedt H, Kurzbeck S. Proc 5th Eur Rheol Conf, Protoroz/SLO 1998:41±4.
[108] Gabriel C, Muendstedt H. Rheol Acta 1999;38:393±403.
[109] Schedenig T, Schausberger A. In: Proceedings of 13th International Conference on Rheology, Cambridge, 2000,
1-263-5.
[110] Winter HH, Chambon F. J Rheol 1986;30:367±82.
[111] Choudary V, Varma HS, Varma KI. Polymer 1991;32:2534±40.
[112] Choudary V, Varma HS, Varma KI. Polymer 1991;32:2541±5.
[113] Coran AV, Patel RB. In: Holden YG, Legge NR, Quirk R, Schroeder HE, editors. Thermoplastic elastomers, 2nd ed.
Munich: Hanser, 1996. p. 154±90.
[114] Neissl W, Ledwinka H. Kunststoffe/plast Europe 1993;33:228±35.
[115] Bernreitner K, Hammerschmid K. In: Karger-Kocsis J, editor. Polypropylene Ð an A to Z reference. Dordrecht: Kluwer
Academic Publisher, 1999. p. 148±57.
[116] Paulik C, Neissl W. Proc ANTEC '98, Soc Plastics Engineers, 1998. p. 2565±9.
[117] Gregor-Svetec D. J Appl Polym Sci 2000;75:1211±20.
[118] Meller M, Luciani A, MaÊnson J-AE. Int Polym Proc 1999;14:221±7.
[119] Ishikawa M, Ushudi Y, Kondo K, Hatada S. Polymer 1996;37:5375±81.
[120] Pople JA, Mitchell GR, Sutton SJ, Vaughan AS, Chai CK. Polymer 1999;40:2769±77.
[121] Riley DW. Plastics Engng 1999/06:49±52.
M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944 943

[122] Huneault MA, Mighri F, Ko GH. Watanabe F. In: Proceedings of ANTEC 2000, Society of Plastics Engineers, 2000,
#0019 (CD-ROM Edition).
[123] Katz HS, Milewski JV, editors. Handbook of ®llers for plastics. New York: Van Nostrand Reinhold, 1987.
[124] Milewski JV, Katz HS, editors. Handbook of reinforcements for plastics. New York: Van Nostrand Reinhold, 1987.
[125] Rothon RN, editor. Particulate-®lled polymer composites. Harlow: Longman, 1995.
[126] Goel CD. Polym Engng Sci 1980;20:198±201.
[127] Rochette A, Choplin L, Tanguy PA. Polym Compos 1988;9:419±25.
[128] Todd DB. Adv Polym Technol 2000;19:54±64.
[129] Bories M, Huneault MA, La¯eur PG. In: Proceedings of PPS 15, 's Hertogenbosch 1999, #277 (CD-ROM Edition).
[130] Han CD. J Appl Polym Sci 1974;18:821±9.
[131] Nanguneri SR, Rao NS, Subramanian N. Rheol Acta 1987;26:301±7.
[132] Kleinecke K-D. Rheol Acta 1988;27:150±61.
[133] Adams JM, Walters K. Macromol Chem: Macromol Symp 1993;68:227±44.
[134] Gahleitner M, Bernreitner K, Neissl W. J Appl Polym Sci 1994;53:283±9.
[135] Suetsugu Y, White JL. J Appl Polym Sci 1983;28:1481±501.
[136] Inn W-Y, Wang SQ. Rheol Acta 1993;32:581±8.
[137] Gahleitner M, Bernreitner K, Knogler B, Neissl W. Macromol Symp 1996;108:127±36.
[138] Wolfschwenger J, Hauer A, Neissl W, Gahleitner M. Proc Euro®llers 1997;'97:375±7.
[139] Hoffmann B, Dietrich C, Thomann R, Friedrich C, Muehlhaupt R. Macromol Rapid Commun 2000;21:57±61.
[140] Steinmann S, FahrlaÈnder M, Fraessdorf W, Friedrich C. In: Proceedings of 13th International Conference on Rheology,
Cambridge, 2000, 1-197-9.
[141] Boutahar K, Carrot C, Guillet J. J Appl Polym Sci 1996;60:103±14.
[142] Pogodina NV, Winter HH. Macromolecules 1998;31:8164±72.
[143] Pogodina NV, Siddiquee SK, Van Egmond JW, Winter HH. Macromolecules 1999;32:1167±74.
[144] Kumaraswamy G, Issaian AM, Korn®eld JA. Macromolecules 1999;32:7537±47.
[145] Kumaraswamy G, Verma RK, Issaian AM, Wang P, Korn®eld JA, Yeh F, Hsiao BS, Olley RH. Polymer 2000;41:8931±
40.
[146] Wassner E, Maier R-D. In: Proceedings of 13th International Conference on Rheology, Cambridge, 2000, 1-183-5.
[147] Bove l, Alfonso GC, Nobile MR. In: Proceedings of 13th International Conference on Rheology, Cambridge, 2000,
1-243-4.
[148] Utracki LA. Polymer alloys and blends. Munich: Hanser, 1989.
[149] Utracki LA. J Rheol 1991;35:1615±37.
[150] Vinckier I, Moldenaers P, Mewis J. J Rheol 1997;41:705±18.
[151] Vinckier I, Mewis J, Moldenaers P. Rheol Acta 1997;36:513±23.
[152] Vinckier I, Moldenaers P, Mewis J. Rheol Acta 1999;38:65±72.
[153] Danesi S, Porter RS. Polymer 1978;19:448±57.
[154] Kim BK, Do IH. J Appl Polym Sci 1996;61:439±47.
[155] D'Orazio L, Mancarella C, Martuscelli E, Polato F. Polymer 1991;32:1186±94.
[156] Yu TC. Proc ANTEC '95, Soc Plastics Engineers, 1995. p. 2347±58.
[157] Bernreitner K, Gahleitner M, Paulik C, Neissl W. Proc PPS 12, Sorrento, 1996. p. 235±6.
[158] Kwang H, Rana D, Cho K, Rhee J, Woo T, Lee BH, Choe S. Polym Engng Sci 2000;40:1672±81.
[159] Graebling D, Muller R. Colloid Polym Sci 1991;55:89±103.
[160] Gramespacher H, Meissner J. J Rheol 1992;36:1127±41.
[161] D'Orazio L, Mancarella C, Martuscelli E, Cecchin G, Corrieri R. Polymer 1999;40:2745±57.
[162] Poelt P, Ingolic E, Gahleitner M, Bernreitner K, Geymayer W. J Appl Polym Sci 2000;78:1152±61.
[163] Plochocki AP. Trans Soc Rheol 1976;20:311±7.
[164] Wu S. Polymer 1985;26:1855±63.
[165] Starke JU, Michler GH, Grellmann W, Seidler S, Gahleitner M, Fiebig J, Nezbedova E. Polymer 1998;39:75±82.
[166] Blom HP, The JW, Rudin A. J Appl Polym Sci 1995;58:995±1006.
[167] Petrovic Z, Budinski-Simendic J, Divjakovic V, Skribic Z. J Appl Polym Sci 1996;59:301±10.
[168] Kim BK, Do IH. J Appl Polym Sci 1996;70:2207±18.
[169] Westphal SP, Ling MTK, Woo L. Proc ANTEC '96, Soc Plastics Engineers, 1996. p. 1629±33.
944 M. Gahleitner / Prog. Polym. Sci. 26 (2001) 895±944

[170] Da Silva ALN, Rocha MCG, Coutinho FMB, Bretas R, Scuracchio C. J Appl Polym Sci 2000;75:692±704.
[171] Suhm J, Schneider MJ, Muehlhaupt R. In: Karger-Kocsis J, editor. Polypropylene Ð an A to Z reference. Dordrecht:
Kluwer Academic Publisher, 1999. p. 104±15.
[172] Paulik C. Diploma thesis, J. Kepler University, Linz, 1992.
[173] Kim G-M, Michler GH, Gahleitner M, Fiebig J. J Appl Polym Sci 1996;60:1391±403.
[174] Gahleitner M, Paulik C, Bernreitner K, Neissl W. Proc Joint Conf Italian, Austrian, Slowenian Rheologists, Rheotech,
Trieste, 1998. p. 69±73.
[175] Liang BR, White JL, Spruiell JE, Goswani BC. J Appl Polym Sci 1983;28:2011±22.
[176] Scholz P, Froehlich D, Muller R. J Rheol 1989;33:481±99.
[177] Pallerne JF. Rheol Acta 1990;29:204±14.
[178] Ohlsson B, Hassander H, Tornell B. Polymer 1998;39:6705±14.
[179] Brahimi B, Ait-Kadi A, Ajji A, Jerome B, Fayt K. J Rheol 1991;35:1069±78.
[180] Willemse RC, Raymaker ESJ, Van Dam J, De Boer AP. Polymer 1999;40:6651±9.
[181] Wang X, Crawford RJ, Fatnes AM, Harkin-Jones E. In: Proceedings of ANTEC 2000, Society of Plastics Engineers,
2000, #0360 (CD-ROM Edition).

You might also like