Sie sind auf Seite 1von 446

Filled

Polymers
Science and
Industrial
Applications
Filled
Polymers
Science and
Industrial
Applications

Jean L. Leblanc

Boca Raton London New York

CRC Press is an imprint of the


Taylor & Francis Group, an informa business
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2010 by Taylor and Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works

Printed in the United States of America on acid-free paper


10 9 8 7 6 5 4 3 2 1

International Standard Book Number: 978-1-4398-0042-3 (Hardback)

This book contains information obtained from authentic and highly regarded sources. Reasonable
efforts have been made to publish reliable data and information, but the author and publisher cannot
assume responsibility for the validity of all materials or the consequences of their use. The authors and
publishers have attempted to trace the copyright holders of all material reproduced in this publication
and apologize to copyright holders if permission to publish in this form has not been obtained. If any
copyright material has not been acknowledged please write and let us know so we may rectify in any
future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced,
transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or
hereafter invented, including photocopying, microfilming, and recording, or in any information stor-
age or retrieval system, without written permission from the publishers.

For permission to photocopy or use material electronically from this work, please access www.copy-
right.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222
Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that pro-
vides licenses and registration for a variety of users. For organizations that have been granted a pho-
tocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are
used only for identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents

Preface.......................................................................................................................xi
Author Bio...............................................................................................................xv

  1. Introduction....................................................................................................1
1.1. Scope of the Book...................................................................................1
1.2. Filled Polymers vs. Polymer Nanocomposites...................................3
References........................................................................................................8

  2. Types of Fillers............................................................................................ 11

  3. Concept of Reinforcement......................................................................... 15
Reference........................................................................................................ 19

  4. Typical Fillers for Polymers...................................................................... 21


4.1 Carbon Black......................................................................................... 21
4.1.1 Usages of Carbon Blacks.......................................................... 21
4.1.2 Carbon Black Fabrication Processes....................................... 21
4.1.3 Structural Aspects and Characterization
of Carbon Blacks....................................................................... 24
4.1.4 Carbon Black Aggregates as Mass Fractal Objects..............30
4.1.5 Surface Energy Aspects of Carbon Black..............................44
4.2 White Fillers.......................................................................................... 49
4.2.1 A Few Typical White Fillers.................................................... 49
4.2.1.1 Silicates......................................................................... 49
4.2.1.2 Natural Silica............................................................... 52
4.2.1.3 Synthetic Silica............................................................ 53
4.2.1.4 Carbonates...................................................................54
4.2.1.5 Miscellaneous Mineral Fillers................................... 56
4.2.2. Silica Fabrication Processes..................................................... 56
4.2.2.1 Fumed Silica................................................................ 56
4.2.2.2 Precipitated Silica....................................................... 58
4.2.3 Characterization and Structural Aspects of
Synthetic Silica.......................................................................... 62
4.2.4 Surface Energy Aspects of Silica............................................ 68
4.3 Short Synthetic Fibers.......................................................................... 69
4.4 Short Fibers of Natural Origin........................................................... 72
References...................................................................................................... 79

v
vi Contents

Appendix 4.................................................................................................... 82
A4.1 Carbon Black Data............................................................................ 82
A4.1.1 Source of Data for Table 4.5............................................... 82
A4.1.2 Relationships between Carbon Black
Characterization Data........................................................84
A4.2  Medalia’s Floc Simulation for Carbon Black Aggregate.............85
A4.3  Medalia’s Aggregate Morphology Approach............................... 86
A4.4  Carbon Black: Number of Particles/Aggregate............................ 89

  5. Polymers and Carbon Black...................................................................... 91


5.1 Elastomers and Carbon Black (CB).................................................... 91
5.1.1 Generalities................................................................................ 91
5.1.2 Effects of Carbon Black on Rheological Properties............. 95
5.1.3 Concept of Bound Rubber (BdR).......................................... 108
5.1.4 Bound Rubber at the Origin of Singular Flow
Properties of Rubber Compounds.......................... ............... 112
5.1.5 Factors Affecting Bound Rubber.......................................... 114
5.1.6 Viscosity and Carbon Black Level........................................ 121
5.1.7 Effect of Carbon Black on Mechanical Properties.............. 125
5.1.8 Effect of Carbon Black on Dynamic Properties.................. 140
5.1.8.1 Variation of Dynamic Moduli with Strain
Amplitude (at Constant Frequency and
Temperature)............................................................. 141
5.1.8.2 Variation of tan δ with Strain Amplitude and
Temperature (at Constant Frequency)...................142
5.1.8.3 Variation of Dynamic Moduli with
Temperature (at Constant Frequency and
Strain Amplitude)..................................................... 142
5.1.8.4 Effect of Carbon Black Type on G′
and tan δ.................................................................... 144
5.1.8.5 Effect of Carbon Black Dispersion on
Dynamic Properties................................................. 146
5.1.9 Origin of Rubber Reinforcement by
Carbon Black............................................................................ 148
5.1.10 Dynamic Stress Softening Effect.......................................... 151
5.1.10.1 Physical Considerations........................................... 151
5.1.10.2 Modeling Dynamic Stress Softening as a
“Filler Network” Effect............................................ 152
5.1.10.3 Modeling Dynamic Stress Softening as a
“Filler–Polymer Network” Effect........................... 168
5.2 Thermoplastics and Carbon Black................................................... 172
5.2.1 Generalities.............................................................................. 172
5.2.2 Effect of Carbon Black on Rheological Properties of
Thermoplastics........................................................................ 173
Contents vii

5.2.3 Effect of Carbon Black on Electrical Conductivity of


Thermoplastics........................................................................ 175
References.................................................................................................... 179
Appendix 5.................................................................................................. 185
A5.1 Network Junction Theory.............................................................. 185
A5.1.1 Developing the Model...................................................... 185
A5.1.2 Typical Calculations with the Network
Junction Model.................................................................. 188
A5.1.3 Strain Amplification Factor from the Network
Junction Theory................................................................. 190
A5.1.3.1 Modeling the Elastic Behavior
of a Rubber Layer between Two
Rigid Spheres................................................... 190
A5.1.3.2 Experimental Results vs.
Calculated Data................................................ 191
A5.1.3.3 Comparing the Theoretical Model with
the Approximate Fitted Equation.. .. ............ 192
A5.1.3.4 Strain Amplification Factor............................ 193
A5.1.4 Comparing the Network Junction Strain
Amplification Factor with Experimental Data............. 194
A5.2 Kraus Deagglomeration–Reagglomeration Model for
Dynamic Strain Softening............................................................. 196
A5.2.1 Soft Spheres Interactions................................................. 196
A5.2.2 Modeling G′ vs. γ0............................................................. 197
A5.2.3 Modeling G″ vs. γ0............................................................ 198
A5.2.4 Modeling tan δ vs. γ0........................................................ 200
A5.2.5 Complex Modulus G* vs. γ0............................................. 202
A5.2.6 A Few Mathematical Aspects of the
Kraus Model...................................................................... 204
A5.2.7 Fitting Model to Experimental Data.............................. 206
A5.2.7.1 Modeling G′ vs. Strain.................................... 207
A5.2.7.2 Modeling G″ vs. Strain.................................... 209
A5.3 Ulmer Modification of the Kraus Model for Dynamic
Strain Softening: Fitting the Model.............................................. 212
A5.3.1 Modeling G′ vs. Strain (same as Kraus)......................... 213
A5.3.2 Modeling G′′ vs. Strain..................................................... 215
A5.4 Aggregates Flocculation/Entanglement Model
(Cluster–Cluster Aggregation Model, Klüppel et al.)............... 218
A5.4.1 Mechanically Effective Solid Fraction
of Aggregate...................................................................... 219
A5.4.2 Modulus as Function of Filler Volume Fraction........... 220
A5.4.3 Strain Dependence of Storage Modulus........................ 221
A5.5 Lion et al. Model for Dynamic Strain Softening........................222
A5.5.1 Fractional Linear Solid Model.........................................222
viii Contents

A5.5.2 Modeling the Dynamic Strain Softening Effect...........223


A5.5.3 A Few Mathematical Aspects of the Model.................. 226
A5.6 Maier and Göritz Model for Dynamic Strain Softening........... 227
A5.6.1 Developing the Model...................................................... 227
A5.6.2 A Few Mathematical Aspects of the Model.................. 229
A5.6.3 Fitting the Model to Experimental Data........................ 230
A5.6.3.1 Modeling G′ vs. Strain.................................... 231
A5.6.3.2 Modeling G″ vs. Strain.................................... 232

  6. Polymers and White Fillers..................................................................... 235


6.1 Elastomers and White Fillers........................................................... 235
6.1.1 Elastomers and Silica.............................................................. 235
6.1.1.1 Generalities................................................................ 235
6.1.1.2 Surface Chemistry of Silica..................................... 236
6.1.1.3 Comparing Carbon Black and (Untreated)
Silica in Diene Elastomers....................................... 237
6.1.1.4 Silanisation of Silica and Reinforcement of
Diene Elastomers...................................................... 239
6.1.1.5 Silica and Polydimethylsiloxane............................. 246
6.1.2 Elastomers and Clays (Kaolins)............................................ 257
6.1.3 Elastomers and Talc................................................................ 260
6.2 Thermoplastics and White Fillers.................................................... 262
6.2.1 Generalities.............................................................................. 262
6.2.2 Typical White Filler Effects and the Concept of
Maximum Volume Fraction.................................................. 266
6.2.3 Thermoplastics and Calcium Carbonates........................... 280
6.2.4 Thermoplastics and Talc........................................................ 291
6.2.5 Thermoplastics and Mica...................................................... 297
6.2.6 Thermoplastics and Clay(s)...................................................300
References.................................................................................................... 302
Appendix 6..................................................................................................308
A6.1 Adsorption Kinetics of Silica on Silicone Polymers...................308
A6.1.1 Effect of Polymer Molecular Weight..............................308
A6.1.2 Effect of Silica Weight Fraction....................................... 310
A6.2 Modeling the Shear Viscosity Function of Filled
Polymer Systems............................................................................. 312
A6.3 Models for the Rheology of Suspensions of Rigid Particles,
Involving the Maximum Packing Fraction Φm........................... 315
A6.4 Assessing the Capabilities of Model for the Shear
Viscosity Function of Filled Polymers......................................... 319
A6.4.1 Effect of Filler Fraction..................................................... 320
A6.4.2 Effect of Characteristic Time λ0...................................... 320
A6.4.3 Effect of Yasuda Exponent a............................................ 321
A6.4.4 Effect of Yield Stress σc................................................... 321
Contents ix

A6.4.5 Fitting Experimental Data for Filled


Polymer Systems.............................................................. 322
A6.4.6 Observations on Experimental Data............................ 323
A6.4.7 Extracting and Arranging Shear
Viscosity Data.................................................................. 324
A6.4.8 Fitting the Virgin Polystyrene Data with the
Carreau–Yasuda Model.................................................. 324
A6.4.9 Fitting the Filled Polystyrene Shear Viscosity
Data................................................................................... 326
A6.4.10 Assembling and Analyzing all Results........................ 332
A6.5 Expanding the Krieger–Dougherty Relationship...................... 335

  7. Polymers and Short Fibers...................................................................... 339


7.1 Generalities......................................................................................... 339
7.2 Micromechanic Models for Short Fibers-Filled Polymer
Composites..........................................................................................344
7.2.1 Minimum Fiber Length.........................................................344
7.2.2 Halpin–Tsai Equations...........................................................345
7.2.3 Mori–Tanaka’s Averaging Hypothesis and Derived
Models...................................................................................... 351
7.2.4 Shear Lag Models.................................................................... 353
7.3 Thermoplastics and Short Glass Fibers........................................... 358
7.4 Typical Rheological Aspect of Short Fiber-Filled
Thermoplastic Melts.......................................................................... 368
7.5 Thermoplastics and Short Fibers of Natural Origin..................... 370
7.6 Elastomers and Short Fibers............................................................. 375
References.................................................................................................... 383
Appendix 7.................................................................................................. 389
A7.1 Short Fiber-Reinforced Composites: Minimum Fiber
Aspect Ratio..................................................................................... 389
A7.1.1 Effect of Volume Fraction on Effective
Fiber Length...................................................................... 389
A7.1.2 Effect of Matrix Modulus on Effective
Fiber Length...................................................................... 390
A7.1.3 Effect of Fiber-to-Matrix Modulus Ratio on
Effective Fiber Length/Diameter Ratio......................... 391
A7.2 Halpin–Tsai Equations for Short Fibers Filled Systems:
Numerical Illustration.................................................................... 391
A7.2.1 Longitudinal (Tensile) Modulus E11............................... 392
A7.2.2 Transversal (Tensile) Modulus E22. ................................ 393
A7.2.3 Shear Modulus G12............................................................ 393
A7.2.4 Modulus for Random Fiber Orientation........................ 394
A7.2.5 Fiber Orientation as an Adjustable
Parameter. ......................................................................................394
x Contents

A7.2.6 Average Orientation Parameters from


Halpin–Tsai Equations for Short Fibers
Filled Systems.................................................................... 394
A7.2.6.1 Longitudinal (Tensile) Modulus E11.............. 395
A7.2.6.2 Transversal (Tensile) Modulus E22. ............... 396
A7.2.6.3 Orientation Parameter X................................. 396
A7.3 Nielsen Modification of Halpin–Tsai Equations with
Respect to the Maximum Packing Fraction: Numerical
Illustration........................................................................................ 396
A7.3.1 Maximum Packing Functions......................................... 397
A7.3.2 Longitudinal (Tensile) Modulus E11............................... 398
A7.3.3 Transverse (Tensile) Modulus Ey.................................... 398
A7.3.4 Shear Modulus G.............................................................. 398
A7.4 Mori–Tanaka’s Average Stress Concept: Tandon–Weng
Expressions for Randomly Distributed Ellipsoidal
(Fiber-Like) Particles: Numerical Illustration............................. 399
A7.4.1 Eshelby’s Tensor (Depending on Matrix Poisson’s
Ratio and Fibers Aspect Ratio Only).............................. 399
A7.4.2 Materials’ Constants (i.e., Not Depending on Fiber
Volume Fraction)...............................................................400
A7.4.3 Materials and Volume Fraction Depending
Constants............................................................................ 401
A7.4.4 Calculating the Longitudinal
(Tensile) Modulus E11. ...................................................... 402
A7.4.5 Calculating the Transverse (Tensile) Modulus E22....... 402
A7.4.6 Calculating the (In-Plane) Shear Modulus G12............. 403
A7.4.7 Calculating the (Out-Plane) Shear Modulus G23..........404
A7.4.8 Comparing with Experimental Data.............................404
A7.4.9 Tandon–Weng Expressions for Randomly
Distributed Spherical Particles:
Numerical illustration...................................................... 406
A7.4.9.1 Eshelby’s Tensor (Depending on Matrix
Poisson’s Ratio Only)....................................... 406
A7.4.9.2 Materials’ Constants (i.e., Not
Depending on Filler Volume Fraction)......... 406
A7.4.9.3 Materials and Volume Fraction
Depending Constants..................................... 407
A7.4.9.4 Calculating the Tensile Modulus E...............408
A7.4.9.5 Calculating the Shear Modulus G.................408
A7.5 Shear Lag Model: Numerical illustration.................................... 409

Index........................................................................................................... 411
Preface

This book is an outgrowth of a course I have taught for several years to


master and doctorate students in polymer science and engineering at the
Université Pierre et Marie Curie (Paris, France). It is also based on around 30
years of interest, research and engineering activities in the fascinating field
of so-called complex polymer systems, i.e., heterogeneous polymer based
materials with strong interactions between phases. Obviously, rubber com-
pounds and filled thermoplastics belong to such systems. If one considers
that, worldwide, around 40% of all thermoplastics and 90% of elastomers
are used as more or less complicated formulations with so-called fillers, it
­follows that approximately 100 million tons/year of polymers are indeed
“filled systems.” Quite a number of highly sophisticated applications of
polymers would simply be impossible without the enhancement of some of
their properties imparted by the addition of fine mineral particles or by short
fibers, of synthetic or natural origin.
The idea that, if a single available material cannot fulfill a set of desired
properties, then a mixture or a compound of that material with another one
might be satisfactory is likely as old as mankind. Adobe, likely the oldest
building material, is made by blending sand, clay, water and some kind of
fibrous material like straw or sticks, then molding the mixture into bricks
and drying in the sun. It is surely one of the oldest examples of reinforce-
ment of a “plastic” material, moist clay, with natural fibers that was already
in use in the Late Bronze Age, nearly everywhere in the Middle East, North
Africa, South Europe and southwestern North America. In a sense, the basic
principle of reinforcement, i.e., to have a stiffer dispersed material to sup-
port the load transmitted by a softer matrix, is already in the adobe brick.
Therefore, the “discovery” of natural rubber reinforcement by fine powdered
materials, namely carbon black, in the dawn of the twentieth century surely
proceeded from the same idea.
At first, mixing rubber and carbon black was pragmatic engineering, it
gave a better and useful set of properties, and the technique could be some-
what mastered, thanks to side developments, such as the internal mixer. The
very reasons for the reinforcing effect remained unclear for a long time and
the question only started to be seriously considered by the mid ­t wentieth
century. Today, some light has been shed on certain aspects of polymer rein-
forcement, as will be reviewed through the book. But the story is surely
not complete because any progress in the field is strongly connected with
either the availability of appropriate experimental and observation tech-
niques or theoretical views about polymer–filler interactions, or (and most
likely) both.

xi
xii Preface

One of the starting points of my deep interest for filled polymers is the
simple observation that, whilst having different chemical natures, a num-
ber of filled polymers, either thermoplastics or vulcanizable rubbers, exhibit
common singular properties. This aspect will be thoroughly documented
throughout the book but a few basic observations are worth highlighting
here. Let us consider for instance the flow properties of systems that are as
(chemically) different as a compound of high cis-1,4 polybutadiene with a suf-
ficient level of carbon black and a mixture of polyamide 66 with short glass
fibers. They share the same progressive disappearance with increasing filler
content of the low strain (or rate) linear viscoelastic behavior. Regarding the
mechanical properties, the effect of either fine precipitated calcium carbon-
ate particles or short glass fibers on the tensile and flexural moduli of poly-
propylene are qualitatively similar but by no means corresponding to mere
hydrodynamic effects. So, many filled polymer systems are similar in certain
aspects and different in others. Understanding why is likely to be the source
of promising scientific and engineering developments.
The possibilities offered by combining one (or several) polymer(s) with one
(or several) foreign stiffer component(s) are infinite and the just emerging
nanocomposites science is an expected development of the science and tech-
nology of filled polymers, once the basic relationships between reinforce-
ment and particle size had been established. For reasons that are given in
Chapter 1, nanofillers have been excluded from the topics covered by the
book, whose objectives are to survey quite a complex field but by no means
offer the whole story.
As stated above, teaching the subject is the origin of the book. In my expe-
rience, nothing must be left in the shadow when teaching a complex sub-
ject and all theories and equations found in the literature must be carefully
checked and weighed, particularly if engineering applications are foreseen.
I am not a theoretician but an experimentalist with an avid interest for any
fundamental approach that might help me to understand what I am measur-
ing. Therefore, whilst theoretical considerations that lead to proposals such
as “property X is proportional to (or a function of) parameter Y,” i.e., X∝ Y or
X∝ F(Y), may be acceptable in term of (scientific) common sense, they are of
very little use for the engineer (and less so for the student) if the coefficient of
proportionality (or the function) is not explicitly given. This is the reason why
all equations displayed in the book have been carefully tested, using (com-
mercial) calculation software. When one loads theoretical equations with
parameters expressed in the appropriate units, then either the unit system
is inconsistent and the software gives no results because the unit equation
is considered, or the right units are used and the results of the theory can be
weighed, at least in terms of “magnitude order.” If the results have the right
order of magnitude, then the theoretical considerations are likely acceptable.
If not… Such an exercise is always useful and I am grateful to my editor
for having accepted, as appendices, a selection of calculation worksheets
(obviously inactive in a printed book) that offer numerical illustrations of
Preface xiii

several of the theoretical considerations discussed in the book. Readers who


are familiar with the calculation software I use will have no difficulties in
implementing these appendices in their own work.
As a last word, it is worth noting that writing a science book on an active
field is (by essence) a never ending task since new interesting contributions
are published every day. But working with an editor forces the scientist-
writer to accept a deadline, in other words to make choices, to develop more
certain subjects and drop other ones, and eventually to bring an end point,
not final but temporary as always in science and industrial applications.

Jean L. Leblanc
Bois-Seigneur-Isaac
Author Bio

Born in 1946, Jean L. Leblanc studied


­physico-chemistry at the University of
Liège, Belgium, with a special emphasis
on polymer science and received his PhD
in 1976, with a thesis on the rheological
properties on SBS bloc copolymers. He then
joined Monsanto Company where, from
1976 to 1987, he held various positions in
the rubber chemicals, the Acrylonitrile-
Butadiene-Styrene plastics (ABS), and the
santoprene• divisions. He left Monsanto in
1987 to join the italian company Montedison
as manager, technical assistance and
applied research, then moved to the position of manager applied research
when Enichem took over Montedison in 1989. In 1988, he became fellow of
the Plastics and Rubber Institute (U.K.) and in 1993 he qualified as European
Chemist (EurChem). In 1993, he was elected Professeur des Universités in France
and joined the Université Pierre et Marie Curie (Paris, France), as head of the
then newly developed polymer rheology and processing laboratory, in col-
laboration with the French Rubber Institute. He is still in this position today
and, since 1997, also teaches polymer rheology and processing at the Free
University of Brussels (Belgium), as a visiting professor. He has written two
books and more than 100 papers.

xv
1
Introduction

1.1  Scope of the Book


This book deals with the properties of filled polymers, i.e. mixtures of
­macromolecular materials with finely divided substances, with respect to
established scientific aspects and industrial developments. So-called (poly-
mer) composites, that consist of long fibers impregnated with resins, such as
glass fibers reinforced polyesters or carbon fibers reinforced epoxy resins, are
not within the subject of this book. Filled polymers discussed hereafter are
heterogeneous systems such that, during processing operations, the polymer
and the dispersed filler flow together. In other words, filled ­polymers are
macroscopically coherent masses that exhibit interesting physical, mechani-
cal, and/or rheological properties, often peculiar, but always resulting from
interactions taking place between a matrix (the polymer) and a dispersed
phase (the filler). It follows obviously that filled polymers have to be prepared
through mixing operations, generally complex and requiring appropriate
machines, in such a manner that a thorough dispersion of filler particles is
achieved.
Why does one prepare filled polymers? There are many reasons, all of
them related to engineering needs. Generally one mixes fillers into polymers
in order to modify properties of the latter, either physical properties, such
as density or conductivity, or mechanical properties, for instance modu-
lus, stiffness, etc., or rheological properties, i.e., viscosity or viscoelasticity.
Occasionally, fillers are also used for economical reasons, as cheap additives
that reduce material costs in polymer applications. Table 1.1 gives the relative
volume costs of a few common mineral fillers in comparison with several
polymers, using polypropylene (PP) as a reference. Clearly, only grinded cal-
cium carbonate and finely divided clays can be considered as “economical”
fillers; in all other cases, specific property improvements are sought when
mixing the filler and the polymer.
A few numbers allow underlining the economical importance of filled poly-
mers. According to recently published market research reports (2007), the
worldwide consumption of fillers is more than 50 million tons with a global
value of approximately €25 billion. Many application areas are concerned,

1
2 Filled Polymers

Table 1.1
Relative Cost of Mineral Fillers and Polymers
Relative Weight Cost
Type of Filler or Polymer (Polypropylene = 1.0)
Grinded calcium carbonate 0.3–0.6
Grinded clays 0.4–0.7
Polyvinyl chloride 0.7
Carbon black 0.7–1.2
Polypropylene 1
Talc 1.1–1.4
Polyethylene 1.1
Calcined clays 1.5–1.7
Wollastonite (not treated) 1.6
Natural rubber 1.6
Ethylene-propylene rubber 1.6–1.9
Treated calcined clays 1.7–1.9
Styrene-butadiene rubber 1.7
Silica 1.7–1.9
Precipitated CaCO3 1.9
Polyamides 3.0–6.0
Note: Table assembled using prices and quotations on the
European market during the first sem­ester of 2008.

such as paper, plastics, rubber, paints, and adhesives. Fillers, either synthetic
or of natural origin are produced by more than 700 companies all over the
globe. In Western Europe, 17 millions tons of thermoplastics were consumed
in 2005 with a significant part in association with 1.7 millions tons of mineral
fillers. Polyvinyl chloride (PVC) and polyolefins (polyethylene PE, PP) are
the main markets for mineral fillers, with calcium carbonate CaCO3 account-
ing for more that 80% of the consumption (in volume). In rubber materials,
more that 90% of the applications concern “compounds”, i.e. quite complex
formulations in which fillers are used at around 50% weight (some 30% vol-
ume). The Western Europe consumption of rubbers was 3.79 millions tons
in 2006 (1.28 MioT natural rubber; 2.51 MioT synthetic elastomers) and some
2.25 millions tons carbon black were used in the interim.
Preparing and using filled polymers is consequently a well established
practice in the polymer field, particularly in the rubber industry where the
first use of carbon black as a reinforcing filler can be traced back to the early
twentieth century. There are consequently a number of pragmatic engineer-
ing aspects associated with the preparation, the development and the appli-
cations of filled polymers, not all yet fully understood, despite considerable
progresses over the last 50 years. As usual, scientific investigations on filled
polymer systems started later than empirical engineering (trial-and-error)
Introduction 3

and it is only the recent development of advanced investigation means that


really boosted research and development work in this area, obviously con-
nected with the contemporary physico-chemistry research on interfaces and
interphases.
Polymers, either elastomers or thermoplastics, offer a great variety of
chemical natures, as well as the fillers, but curiously common effects and
properties are (at least qualitatively) observed whatever is the chemistry of
the polymer matrix and of the filler particles. This striking observation is the
very origin of this book that intends to offer a survey of a quite complex field,
with the objectives to highlight what most filler–polymer systems have in
common, how proposed theories and models suit observations and, eventu-
ally what are the specificities of certain filled polymers.

1.2  Filled Polymers vs. Polymer Nanocomposites


A filled polymer system is thus a polymer in which a sufficient quantity (vol-
ume) of a small size foreign rigid (or at least less flexible) material, e.g., pow-
dered minerals, short glass fibers, etc., has been well dispersed in order to
improve certain key properties of engineering importance, for instance mod-
ulus, stiffness, or viscosity. The reinforcing effect of carbon black in rubber is
known for one century (1907, Silvertown, UK) and the mastering and under-
standing of its scientific aspect has led to the development of many high engi-
neering performance products, for instance the automobile, truck, or aircraft
tires. Starting in 1984, a series of patents obtained by Toyota1 described the
use of organoclay additives for plastics as well as various plastic structures
that could replace traditional components (e.g., ­aluminium parts) in automo-
tive applications. Typically U.S. patent No. 4,810,734 described a production
process for a composite material by firstly treating a layered smectite mineral
having a cation exchange capacity (e.g., a phyllosilicate) with a swelling agent
having both an onium ion and a functional group capable of reacting with a
polymer and secondly forming a complex with a molten polymer. U.S. pat-
ent No. 4,889,885 described a composite material made with at least one resin
selected from the group consisting of a vinyl-based polymer, a thermosetting
resin and a rubber, and a layered bentonite uniformly dispersed in the resin.
The layered silicate has a layer thickness of about 0.7–1.2 nm and an interlayer
distance of at least about 3 nm, and at least one polymer ­macromolecule has
to be connected to the layered silicate. Such patents prompted, over the last
20 years, a kind of cult research area for so-called polymer nanocomposites,
whose origin of reinforcement is on the order of nanometers, but with the
capability to deeply affect the final macroscopic properties of the resulting
material. In certain cases, such materials exhibit properties not present in the
pure polymer resin, whilst keeping the processibility, the other mechanical
4 Filled Polymers

polymer properties and the specific weight. Several types of polymeric


nanocomposites can in principle be obtained with different particle nano-
size, nature and shape: clay/polymer, carbon nanotubes, and metal/polymer
nanocomposites.
Let us consider the case of clay/polymer nanocomposites. The key aspect
is obviously the successful formation of suitable clay/polymer nanostruc-
tures, essentially through an intercalation process. In the case of hydrophilic
­polymers (typically polyamides) and silicate layers, pretreatment is not nec-
essary; but most polymers are hydrophobic and are not compatible with
­hydrophilic clays. Complicated and expensive pretreatments are thus required.
For instance organophilic clays can be obtained from normally hydrophilic
clay by using amino acids, organic ammonium salts, or tetra organic phos-
phonium solutions, to name a few reported techniques. Established methods
are: solution induced intercalation, in situ polymerization, and melt process-
ing. Solution induced intercalation consists of solubilizing the polymer in an
organic solvent, then dispersing the clay in the solution and subsequently
either evaporating the solvent or precipitating the polymer. Such a technique
is obviously expensive, raises a number of environmental, health, and safety
problems (common to all organic solution techniques), and in fact leads to
poor clay dispersion. In situ polimerization consists of dispersing clay lay-
ers into a matrix before polimerization, i.e., mixing the silicate layers with
the monomer, in conjunction with the polymerization initiator and/or the
catalyst. This technique is obviously limited to polymers whose ­monomers
are liquids, and therefore excludes most of the general purpose (GP) resins,
namely polyolefins. In the melt processing technique the silicates layers, pre-
viously surface treated with an organo-modifier, are directly dispersed into
the molten polymer, using the appropriate equipment and procedure. A priori,
this technique would be the preferred route with most GP polymers, provid-
ing mixing/dispersion problems are mastered.
In theory, extraordinary improvements of material properties are expected
with polymer nanocomposites but, in reality, the overall balance of usage
properties (i.e., mechanical, hardness, wear resistance, to name a few) in the
best clay/polymer nanocomposites are much lower than in conventional fiber
reinforced composites, or even in certain traditional filled compositions. It is
indeed only in the low filler range, typically 4–5 wt%, and providing the
dispersion of nanoparticles is nearly ideal, that nanocomposites show better
mechanical performances, but at the cost of major difficulties in mass fabrica-
tion. At higher loading, the surface area of the silicate-filler increases, which
leads to insufficient polymer molecules adsorbed on the clay surface. One
may consider that polymer nanocomposites combine two concepts: compos-
ites (i.e., heterogeneous systems) and nanometer-size ­materials; the hope that
manufacturing composites polymer material could eventually be achieved
with a tight control at molecular level (i.e., the nanometer range) surely justi-
fies fundamental research in this area, even if large scale industrial applica-
tions are not yet in sight. Certain thermoplastics, filled with nanometer-size
Introduction 5

materials, have indeed different properties than systems filled with conven-
tional mineral materials. Some of the properties of nanocomposites, such as
increased tensile strength, are routinely achieved by using higher conven-
tional filler loading, but of course at the expense of increased weight and
sometimes with unwanted changes in surface aspects, i.e., gloss with certain
polymers. Obviously other typical properties of certain polymer nanocom-
posites such as clarity or improved barrier properties cannot be duplicated
by filled resins at any loading.
One may indeed consider that polymer nanostructured materials repre-
sent a radical alternative to the conventional filled polymers and polymer
blends, because the utility of inorganic nanoparticles as additives to enhance
polymer performance has been well established at laboratory level. The
incorporation of low volume (1–5 wt%) of highly anisotropic nanoparticles,
such as layered silicates or carbon nanotubes, results in the enhancement of
certain properties with respect to the neat polymer that are comparable with
what is achieved by conventional loadings (15–40 wt%) of traditional fillers.2
In principle the lower loadings would facilitate processing and reduce com-
ponent weight, and in addition, certain value added properties not normally
possible with traditional fillers are also observed, such as higher stiffness,
reduced permeability, optical clarity, and electrical conductivity. But the
chemical and processing operations to disrupt the low-dimensional crystal-
lites and to achieve uniform distribution of the nanoelement (layered silicate
and single wall carbon nanotube, respectively) continue to be a challenge.
Most commercial interest in nanocomposites has so far focused on ther-
moplastics, essentially because certain polymer nanocomposites allow the
substitution of more expensive engineering resins with less expensive com-
modity polymer nanocomposites, to yield overall cost savings. But such
favorable cases are rare and restricted to very specific applications. A recent
study by a market research company claims that, by 2010, nanocomposites
demand will grow to nearly 150,000 tons, and will rise to over 3 million tons
with a value approaching $15 billion by 2020.3 So far however the market for
these new materials has not developed as expected and if, indeed, exfoliated
(or surface treated) nanoclays are commercially available,4 their uses seem
restricted to very specific cases. Packaging and parts for motor vehicles are
nevertheless expected to be key markets for nanoclay and nanotube compos-
ites. With respect to the improved barrier, strength and conductive proper-
ties that they can offer, polymer nanocomposites should somewhat penetrate
certain food, beverage, and pharmaceutical packaging applications, as well
as specific parts for electronics. In motor vehicles, automotive manufacturers
are expected to consider polymer nanocomposites either as replacement for
higher-priced materials, or to increase the production speed of parts and to
reduce motor vehicle weight by lightening a number of exterior, interior, and
underhood applications. The future will weigh such expectations.
Over the last decades, a considerable number of research papers have been
published whose main subject is so-called polymer nanocomposites,5 i.e.,
6 Filled Polymers

mixtures or preparations involving macromolecular materials and small


particles with dimensions in the nanometer range, with however a great deal
of confusion in the author’s opinion. Indeed, a careful reading of published
papers reveals that for certain authors, nanoparticles are entities with (equiv-
alent) diameters up to a few tens nanometers, whilst others title their works
with the heading nanocomposites but consider mixtures with particles in
the micron range. It is also worth underlining that nanoparticles technol-
ogy implies that individual representatives particles (i.e., spheres, platelets,
etc.) are ideally dispersed in the polymer matrix, without agglomeration or
flocculation. This aspect of polymer nanocomposites appears thus in sharp
contrast with conventional filled polymer technology where elementary
particles must be suitably clustered in complex tri-dimensional structures
called “aggregates” to yield reinforcing properties. As will be extensively
described in this book, this is the key aspect of the reinforcement of rub-
ber with carbon black and high structure silica. In many published papers
this ideal dispersion of nanoparticles is neither documented nor granted by
the preparation (mixing) process, and therefore the reference to polymer
nanocomposites is dubious. Despite the lack—so far—of significant indus-
trial applications, polymer nanocomposites seem to be a fashion subject for
fundamental research, with sometimes an unfortunate lack of reference to
earlier works on more classical filled polymer systems, namely filled rubber
materials, surely the oldest class of complex polymer materials of industrial
importance. There are a number of recent books, reviews, and treatises on
so-called polymer nanocomposites6–8 and elastomer nanocomposites.9,10 The pres-
ent book is definitely not addressing the same subject, but rather so-called
“filled polymer systems” that are nowadays used yearly in quantities of hun-
dred thousands to million tons worldwide.
In order to avoid confusion it is thus necessary to clearly define what are
filled polymer systems, the very subject of the present book, in contrast with
polymer naonocomposites. It is clear that industrial use is not a sufficient
criterion to distinguish both classes of materials. Whilst mainly concerned
with rubber reinforcement, Hamed offered recently quite a clear and well-
supported proposal to distinguish filled polymer systems, with respect to the
smallest size d of the dispersed phase.11 The characteristic smallest dimension
d depends of course of the actual geometry of the particles, the diameter for
spheres and rods, the thickness for plates and scales. There are a number of
available materials whose characteristic particle dimension is in the 1–100 nm
range and therefore the prefix nano is ambiguously used in the literature. We
will consequently somewhat follow the Hamed’s proposal: when the charac-
teristic dimension d of the dispersed phase is between 1 and 10 nm, then one
is dealing with nanocomposites, when 100 nm > d > 10 nm, then mesocom-
posites are involved, with d above 100, composite materials are referred with
the prefix micro, and the prefix macro when very gross  (d > 104  nm) rigid
“entities” are dispersed in a polymer. Further to this basic characterization,
Hamed considers that the dispersed entities can be structured, either a priori
Introduction 7

by their nature or through their manufacturing process, or as a result of the


kinetics and thermodynamics of phase separation that may occur during the
preparation of the complex polymer system. The proposal is further elabo-
rated in Table 1.2., with typical examples of concerned materials.
With respect to Table 1.2, all filled polymer systems discussed in this book
are either meso or microcomposites, and most of them have a considerable
industrial importance. The proposal by Hamed is based on well sounded
arguments on the mechanical properties of filled rubbers and is further
reinforced by very recent observations on the likely origin of the unusual
properties of (true) nanocomposites. Indeed as demonstrated by a number
of authors, so-called anomalous rheological and mechanical properties of
polymer nanocomposite systems are observed when the characteristic
dimensions of (ideally) dispersed particles are in the 1–10 nm range, in fact
commensurable with some typical dimensions of polymer dynamics, namely
the reptational tube diameter (a few nanometers), as considered when mod-
eling the behavior of entangled polymers. In fact polymer nanocomposites
are distinguished by the convergence of length scales corresponding to the
radius of gyration of the polymer chains, a dimension of the nanoparticle
and the mean distance between the nanoparticles.12 It was therefore hypoth-
esized that, when nanoparticles have such small dimensions, they have the
capability to participate in the local polymer dynamics.13
Filled polymer systems of industrial importance, e.g., filled rubber com-
pounds, filled thermoplastics are thus meso or microcomposites, ­possibly
with a structuration (of the dispersed phase) at the nano or meso scale.
Whilst no sizeable commercial application yet exist for nanocomposites rub-
bers or thermoplastics (to the author’s knowledge), considerable research has
been made since 1984 with so-called ex-foliated layered silicate “nano-clays.”
Exfoliation means that individual clay sheets, of around 1 nm thickness,
have been separated and adequately dispersed in the matrix. Some rein-
forcement has indeed been demonstrated with such exfoliated nanoparticles
but, generally with very specific rubber systems and/or at a cost of prepara-
tion that is hardly compatible with reasonable chances of commercialization.

Table 1.2
Classification of Filled/Composite Polymer Systems
Characteristic
Designation Dimension (nm) Example
Nanofiller/particle composite 1–10 Polyamide/exfoliated montmorillonite
Rubber compounds with highly
Mesofiller/particle composite 10–100
reinforcing carbon blacks
Polypropylene/grinded calcium
Microfiller/particle composite 100–10,000
carbonate
Macrofiller/particle composite  > 104 Polymer concrete
8 Filled Polymers

It can further be commented that the level of reinforcement obtained in such


systems is not even comparable with what is practically achieved with con-
ventionally filled mesocomposite polymers, namely rubbers. No amorphous
vulcanized rubber reinforced only with exfoliated clay has been reported to
have a tensile strength in the 30 MPa range, as currently obtained with con-
ventional carbon black filled compounds. One can nevertheless expect that,
owing to their special geometries (plates or scales), properly dispersed exfo-
liated clays might enhance certain properties, such as gas impermeability,
through barrier effects, or thermal or electrical conductivity, through appro-
priate orientation effects, and therefore find niche markets.

References
1. U.S. Patents: 4,472,538 (Composite material composed of clay mineral and
organic high polymer and method for producing the same, September
18, 1984); 4,739,007 (Composite material and process for manufacturing
same, April 19, 1988); 4,810,734 (Process for producing composite material,
March 7, 1989); 4,889,885 (Composite material containing a layered silicate,
December 26, 1989); 5,091,462 (Thermoplastic resin composition, February
25, 1992).
2. Q. Yuan, R.D.K Misra. Polymer nanocomposites: current understanding and
issues. Mater. Sci. Technol., 22 (7), 742–755, 2006.
3. Nanocomposites. The Freedonia Group, Inc., Cleveland, OH, 2006.
4. For example, Nanomer® nanoclays from AMCOL Intern. Corp., Arlington
Heights, IL; Cloisite® and Nanofil® from Southern Clay Products, Inc.,
Gonzales, TX; Bentone® from Elementis plc, Hightstown, NJ.
5. See for instance the following recent reviews: S.S. Ray, M. Okamoto. Polymer/
layered silicate nanocomposites: a review from preparation to processing. Prog.
Polym. Sci., 28 (11), 1539–1641, 2003; H. Fischer. Polymer nanocomposites: from
fundamental research to specific applications. Mater. Sci. Eng. C, 23 (6–8), 763–
772, 2003; Wang, Z.-X. Guo, S. Fu, W. Wu, D. Zhu. Polymers containing fuller-
ene or carbon nanotube structures. Prog. Polym. Sci., 29 (11), 1079–1141, 2004;
J. Jordan, K.I. Jacob, R. Tannenbaum, M.A. Sharaf, I. Jasiuk. Experimental trends
in polymer nanocomposites—a review. Mater. Sci. Eng. A, 393 (1–2), 1–11, 2005.
6. P.M. Ajayan, L.S. Schadler, P.V. Braun. Nanocomposite Science and Technology.
Wiley, New York, NY, 2003. ISBN: 9783527303595.
7. Y.-W. Mai, Z.-Z. Yu Ed. Polymer Nanocomposites. CRC Press, Baton Roca, FL,
USA; 2006. ISBN 9780849392979; a review by an international team of authors
with 13 papers on layered silicates/polymer compositions and eight papers on
nanotubes, nanoparticles and inorganic-organic hybrid systems.
8. J.H. Koo. Polymer Nanocomposites. McGraw-Hill Prof., New York, NY, 2006. ISBN
13: 978-0071458214.
Introduction 9

9. S.D. Sadhu, M. Maiti, A.K. Bhowmick. Elastomer-clay nanocomposites.


Chapter  2, 23–56. In Current Topics in Elastomer Research, A.K. Bhowmick Ed.
CRC Press, Taylor & Francis Group, Boca Raton, FL, 23–562008. ISBN-13: 978-0-
8493-7317-6.
10. M. Maiti, M. Bhattacharya, A.K. Bhowmick. Elastomer nonocomposites. Rubb.
Chem. Technol., 81, 384–469, 2008.
11. G.R. Hamed. Rubber reinforcement and its classification. Rubb. Chem. Technol.,
80, 533–544, 2007.
12. R. Krishnamoorti , R.A.Vaia. Polymer nanocomposites. J. Polym. Sci. Part B.
Polym. Phys., 45 (24), 3252–3256, 2007.
13. M.E. Mackay. Anomalous rheology of polymer-nanoparticle suspensions.
XVth International Congress on Rheology, Monterey, CA, August 3–8, 2008. Paper
KL.11.
2
Types of Fillers

In polymer technology, there are essentially two major classes of fillers,


either extracted or fabricated. Minerals such as talc and clays (Al2O3, 2SiO2,
2H2O) are extracted, grinded, and possibly treated and therefore belong to
the first class. Calcite (CaCO3) belongs to both classes, as it can be either
extracted and grinded or obtained through a chemical process that involves
precipitation. Carbon blacks result from the incomplete combustion of
hydrocarbon feedstock, and are consequently fabricated fillers, as well as
synthetic silica that are obtained through more or less complex chemical
operations. Short fibers made either of glass, or of carbon, are fabricated
products, and we arbitrarily include cellulose fibers also in the second class,
because quite complex treatments are required before they can be used as
a polymer reinforcing material. Moreover, many types of natural fibre have
been considered for use in polymers as reinforcing materials including
flax, hemp, jute, straw, wood flour, rice husks, sisal, raffia, green coconut,
banana, and pineapple leaf fibre to name a few, but technical problems such
as moisture absorption and low impact strength have sometimes restricted
their development. Wood flour nowadays used to prepare so-called wood–
polymer composites (WPC), which represents a growing market over the
last decades,* can also be considered as a fabricated filler with respect to its
preparation mode.
Fillers for polymers exhibit in fact a stunning variety of chemical natures,
particle sizes and shapes. Essentially three basic shapes can be distinguished:
either spheres, or plaques (disks, lamellas) or rods (needles, fibers), as illus-
trated in Figure 2.1. Such basic shapes can be further combined to result in
quite complex geometrical objects to which specific (reinforcing) properties
can be associated. Carbon black aggregates offer typical examples of com-
plex tri-dimensional structures whose shape specifically affects the reinforc-
ing properties, as will be discussed hereafter. Most fillers, either extracted or
fabricated, have a mineral origin, with the notable exception of carbon blacks
that result from the thermal degradation of hydrocarbons. There are also a

* In North America the WPC market amounts today to around 300,000 tons/year, essen-
tially for building and garden applications, particularly decking and associated prod-
ucts. Estimated over $600 Mio in 2002, the USA and Canada segment is nowadays worth
over $2 billion and worldwide estimates are in the $3 billion range. Market growth is
slower in West Europe with a consumption of around 140,000 tons in 2002, over 200,000
tons in 2005 and estimated to reach some 270,000 tons in 2010 (source: A. Eder. WPCs – an
updated worldwide market overview including a short glance at final consumers. 3rd Wood
Fibre Polymer Composites Symposium, Bordeaux, France, March 21–27, 2007).

11
12 Filled Polymers

Spheres Scales, flakes Cylinders, rods,


lamellas needles, fibers

Partial fusion elementary particles => aggregates

Complex tri-dimensional object


=> structural effect of the filler

Figure 2.1
Fillers basic shapes and structure.

number of filler materials that have a vegetal origin, for instance wood flour,
sisal, coco, or jute fibers.
It is tempting to consider a classification scheme for polymer fillers but
no overall system is available and the analysis of existing proposals reveal
that their validity and interest strongly reflect the application considered. We
will nevertheless consider a few logical possibilities, which underline certain
specific aspects of the common property considered.
Considerations based on the refractive index allow to draw a clear dis-
tinction between a filler and a pigment, whilst if certain fillers can be used
to modify the color of a polymer (e.g., carbon black in polyolefin), not all
pigmenting materials have reinforcing capabilities. Let us consider various
materials and their respective refractive indices (Table 2.1). The refractive
index of vacuum is (by definition) equal to 1, and most polymers exhibit
indices around 1.5. One would consider that any given material has no capa-
bility to modify the color of another one if the respective refractive indices
of both materials do not differ by more than 0.2. It follows that materials
with refractive indices either above 1.3 or below 1.7 have practically neither
clearing nor darkening effects on polymers. Consequently, a mineral whose
refractive index is above 1.7 can potentially be used as a pigment (but can
also have reinforcing capabilities), whilst materials whose refractive index is
below 1.7 would be essentially considered as fillers.*
A logical and broader approach would associate the origin, the produc-
tion process and the reinforcing capabilities (Figure 2.2). In this manner,
essentially four types of filler are considered: organic fillers of natural ori-
gin (liege, wood flour, vegetal fibers), organic fillers obtained by chemical

* One notes however that such a classification makes no sense for “dark” fillers, such as carbon
blacks, which do not refract light.
Types of Fillers 13

Table 2.1
Distinguishing Between Filler and Pigment with
Respect to Refractive Index
Material Refractive Index
Vacuum 1.00
Water 1.33
Chalk 1.35
Polymers 1.50
Silica 1.55
BaSO4 1.64
ZnO 2.08
ZrO2 2.17
ZnS 2.37
Diamond 2.42
TiO2, Anatase 2.55
TiO2, Rutile 2.75

Filler

Organic Inorganic

Natural Synthetic Natural Synthetic

- Inactive - Inactive - Inactive - Inactive


- Semiactive - Semiactive - Semiactive - Semiactive
- Active - Active - Active - Active
Liege Synthetic resins Minerals Carbon blacks
Wood flour Cellulose derivatives (CaCO3, talc, Silicas (fumed, precipitated)
Fibres (jute, clays,...) Metal oxides (TiO2, ZnO,...)
sisal,...) Metal salts (BaSO4,...)

Figure 2.2
Classifying fillers with respect to fabrication process and reinforcing activity.

synthesis (synthetic resins, cellulose derivatives), mineral fillers of natural


origin (essentially all extracted fillers) and mineral fillers obtained through
chemical processes in the broad sense (carbon blacks, fumed and precipi-
tated silica). Furthermore, for each type, one might distinguish materials as
active, semiactive, or inert filler, depending how they boost, improve or do
not affect certain mechanical properties of interest, for instance stiffness,
tensile or flexural strength, and abrasion resistance, to name a few.
Another approach, maybe less subjective, consists of paying attention to
particle size because, as illustrated in Figure 2.3, there is a clear relation-
ship between this characteristic and the reinforcing capabilities. Essentially
14 Filled Polymers

105
Grinded CaCO3 mica, talc
Degradative
fillers
104

Dilution Clays
fillers
103
Semireinforcing Precipitated CaCO3
fillers TiO2, ZnO
102
Reinforcing Si aluminates
fillers
Ca silicates
101
Hydrated silica
Particle
size Anhydrous silica
(nm)
Carbon blacks

Figure 2.3
Classifying fillers with respect to particle sizes.

no reinforcement is obtained when particles are larger than 103 nanome-


ter (nm) and too large particles deteriorate mechanical properties of poly-
mer materials. The wide range of particle sizes (and structures) offered
by the manufacturing of carbon blacks and synthetic silica clearly reflect in
the  semireinforcing and reinforcing character of these fillers. The general
relationship between reinforcing capabilities and particle size suggests obvi-
ously that a poorly dispersed mineral, whatever its ultimate particle size, is
likely to deteriorate ultimate mechanical properties, for instance by reducing
the elongation at break of vulcanized rubbers and thermoplastics. Indeed,
large and badly dispersed particles are fracture initiation sites.
3
Concept of Reinforcement

Whilst they can be added to polymers for other purposes, it is mainly for their
reinforcing capabilities that certain fillers offer the largest interest. When
compared to polymers, any mineral exhibits mechanical properties, such
as modulus, stiffness, hardness, that are several order of magnitudes larger.
Therefore, one may reasonably expect that mixing the latter with the former
will result in a heterogeneous mixture that exhibits macroscopic mechanical
properties, at least intermediate between those of the polymer and those of
the filler. Reinforcement of elastomers by carbon black, discovered in 1907
in Silvertown, UK, is likely the most significant example of this effect, that
really permitted the development of the emerging tire technology, strongly
connected of course with the automotive industry.
Essential in rubber technology, the concept of reinforcement is however
very complex, even if relatively easy to capture at first sight. Indeed, when a
filler is added to a polymer, practically all properties are affected, some in a
positive manner, others negatively with respect to a given application. There
has been much debate about which particular property should be considered
as the most expressive in terms of reinforcement. In this respect, it is worth
quoting here the opinion expressed by G. Kraus:1

A precise definition of the term «reinforcement» is difficult because it


depends somewhat on the experimental conditions and the intended
effects of the filler addition…it appears preferable to regard reinforce-
ment broadly as the modification of the viscoelastic and failure proper-
ties of a rubber by a filler to produce one or more favorable results…

The reinforcing capabilities of a filler must consequently be appreciated


with respect to a balance of properties, whose choice depends on the applica-
tion considered. Let us consider the general trends exhibited by a rubber
compound in which increasing quantities of active (e.g., carbon black) or
inert (e.g., finely divided clay) have been added.
As shown in Figure 3.1, certain properties will only either increase or
decrease, for instance viscosity, hardness, but other ones will pass through
extremes in the case of the reinforcing filler. This immediately suggests that
there will be optimum loadings, for a given filler, in a given polymer, for
a specific application. To establish the optimum filler level is therefore the
most important task for the compounder, further complicated by the obvi-
ous requirement that the compound must remain processible at reasonable

15
16

(Mooney) viscosity Hardness Compression set

100 90 50

80 80 Active 40
Active Active

%
60 70 30

Shore A
40 60 20 Inert

ML(1+4) at 100°C
Inert
Inert
20 50 10

Filler level (phr) Filler level (phr) Filler level (phr)

Tensile strength Elongation at break Abrasion

20 1000 500
Active Inert
400
15
Active

%
300

MPa
mm3

10 500
200 Active
5 Inert
Inert 100

Filler level (phr) Filler level (phr) Filler level (phr)

Figure 3.1
Relative variation of rubber compound properties as imparted by active (reinforcing) or inert filler.
Filled Polymers
Concept of Reinforcement 17

energy and labor costs; sometimes the excessive viscosity increase imparted
by very active fillers, either limits their practical level in certain elastomers or
requires additional modification in formulation, for instance higher  levels of
processing oils, or plasticizers, which generally have a penalizing effects on
certain mechanical properties of the vulcanized part.
In general, the reinforcing activity of a filler depends on at least four
criteria:

• The particle size (always smaller than 100 nm)


• The structure (i.e., the spatial organization)
• The specific area
• The surface (chemical) activity.

The structure of the filler material refers to the fact that, during their man-
ufacturing process, reinforcing fillers develop very complex tri-dimensional
shapes, which are called aggregates in the case of carbon black. Aggregate
structure appears thus as one of the most important aspect of reinforcement
and is obviously related with the specific area. The quantification of structure
and the measure of specific area are somewhat related, essentially because
the adsorption of molecules of known size is used to assess both charac-
teristics. The well-known BET (Brunauer, Emmet, Teller) method is used to
measure the adsorption isotherm of nitrogen (N2) absorbed by powdery fill-
ers, whilst the aggregate complexity is assessed by evaluating the maximum
quantity  of larger molecules (for instance di-butylphthalate DBP, or cetyl-
triethylammonium bromide CTAB) than can be adsorbed on the external
surface. As might be expected, there is a (loose) correlation between the so-
called BET surface and the activity (or reinforcing capability) of a filler:

BET < 10 m2/g: inert filler


BET = 10–60 m2/g: semiactive filler
BET > 60 m2/g: active filler
BET > 100 m2/g: very active filler

In fact, relationships between the reinforcing abilities and the character-


istics of the filler are very complicated and, in general, one has to consider
more than one criterion to make valid comparisons, useful for a given filler
in a given polymer, for a given application.
It is worth underlining that the concept of reinforcement has been more
debated in the field of rubber science and technology than in the field of
thermoplastics. The fact that, without suitable reinforcement, most elasto-
mers exhibit so low mechanical properties that no interesting applications
are possible is surely a reason. Another one is that most general purpose
thermoplastics have known their tremendous development in the second
18 Filled Polymers

half of the twentieth century, in parallel with the expansion of petro-


chemistry, and have found immediately interesting applications “as such,”
nearly without additives except a few protective chemicals. Polyethylene
and polypropylene for instance are used to fabricate sheets and films by
essentially exploiting their capabilities as semicrystalline polymers. No
filler is needed to obtain the high mechanical properties that develop when
crystalline structures are properly established and oriented. Polystyrene,
ABS and other styrenics exhibit properties directly used in a number of
applications, without the need of reinforcing fillers. Of course, in their
usages, most thermoplastics must also meet a balance of properties but,
except maybe polyvinyl chloride (PVC), the right material for a given appli-
cation is obtained by controlling the macromolecular size and structure,
essentially through a suitable adaptation of the polymerization process.
The key role played by polymerization catalysts in the developments of
polyolefins clearly supports this point. PVC is an exception because when
suitably compounded with stabilisers, plasticizers, and other ingredients, a
whole range of products can be obtained, essentially by changing the glass
transition temperature of the material. It is quite symptomatic that the so-
called “plastograph,” a small laboratory mixer, was specifically developed
in the 1950s as a convenient tool to document the “plasticization” of PVC.
The addition of fillers to thermoplastics polymers is thus quite a recent
practice, around three decades old, whilst filled rubber compounds are
used for more than a century.
There is nevertheless another important, more technical reason for the
different meaning of reinforcement in the rubber and plastics fields. In
most of their applications, thermoplastics are used within the limits of
their elastic behavior, generally below 10% strain. Indeed, once the yield
strength limit is exceeded, permanent deformation occurs. It follows that
most applications of thermoplastics are first concerned by the elastic behav-
ior of the material; the viscoelastic character plays a secondary role, namely
in what the long term variation of modulus is concerned through the creep
phenomenon for instance. The situation is totally different with rubber
materials, whose performance are controlled by their viscoelastic charac-
ter, in a strain range that is substantially larger than for thermoplastics.
For instance, with rubber materials, the tensile (Young) modulus is far less
significant than the 100 or 200% modulus in most applications. It follows
that the role played by fillers in “reinforcing” rubbers and thermoplastics
is substantially different, as well as the balance of properties, as will be
largely underlined throughout the book. A direct consequence is that the
modeling of the filler’s effects in thermoplastics and in elastomers, whilst
sometimes based on a similar theoretical background, is generally substan-
tially differing in the supporting reasoning and therefore in the applicabil-
ity. It is one of the objectives of this book to identify both the similitudes
and the differences in those theoretical approaches, with respect to the
class of polymer matrix considered.
Concept of Reinforcement 19

Reference
1. G. Kraus. Reinforcement of elastomers by carbon black. Adv. Polym. Sci.,
8, 155–231, 1971.
4
Typical Fillers for Polymers

4.1  Carbon Black


4.1.1 Usages of Carbon Blacks
Essentially, carbon black is the soot that results from the incomplete combus-
tion of hydrocarbon materials, i.e., gas and oils. This definition does not pay
tribute however to the high degree of development and control in use today
in most industrial processes. The uses and the basic production principles
of carbon black are lost in antiquity, but the development of controlled fab-
rication processes dates back to the previous century, resulting nowadays in
highly sophisticated technologies with the capability to produce very fine
and structurally complex materials, in accordance with the most recent stan-
dards of quality.
As we will briefly see below, the term “carbon blacks” covers a very broad
range of filler materials, with numerous applications, as outlined in Table 4.1.
Except elastomer reinforcement, printing inks and several uses in the electri-
cal industry, most application concern relatively low fraction of carbon black,
typically below 5% volume.

4.1.2  Carbon Black Fabrication Processes


Fabrication processes of carbon blacks all share the same principle: con-
trolled heat decomposition of hydrocarbon products. Such processes are
essentially chemical, either thermo-oxidative or mere thermal decomposi-
tion, as described in Table 4.2.
Amongst the thermo-oxidative processes, the furnace black one is the most
recent and nowadays the most important. As illustrated in Figure 4.1, the
liquid combustible (either oil or gas) is sprayed in a flame of natural gas and
hot air. Black smoke is produced that is a mixture of gas and carbon particles,
initially nearly spherical elementary particles that partially fuse together to
produce complex tri-dimensional objects called aggregates. Carbon black
aggregates are quenched through water spraying that stops the pyrolysis
and aggregation processes and cools down smoke, which is then filtered to
recover solid particles. Unburned gas is treated and recycled in the process.
21
22 Filled Polymers

Table 4.1
Important Uses of Carbon Blacks
Domain Application
Elastomers Reinforcing filler in tires and mechanical rubber goods
Printing inks Tinting, rheology modifier
Enduction Black and gray tinting, color enhancement
Thermoplastics Black and gray tinting, color enhancement, anti-UV protection of
polyolefins, high voltage cable shielding, application in
semiconductors, static electricity dissipation
Fibers Tinting
Paper Black and gray tinting, photograph protective paper
Building Cement and concrete tinting
Electrical industry Electrodes, dry batteries and cells

Table 4.2
Fabrication Methods of Carbon Blacks
Chemical Process Method Raw Material (Feedstock)
Thermo-oxidative Furnace black Aromatic oils from coal tar or
decomposition petrol distillates, natural gas
Gas black (Degussa process) Coal tar distillates, natural gas
Lamp black Aromatic oils from coal tar or
petrol distillates
Thermal decomposition Thermal black Natural gas (or oils)
Acetylene black Acetylene

Thermo-oxidative process : furnace black


Liquid feedstock atomized and
Air sprayed into the flame

Smoke treatment for


carbon black recovery

Flame from combustion Smoke gas


of gas combined with Water quench
Gas preheated air (stops pyrolysis)
Oil

Figure 4.1
Carbon black manufacturing process for furnace black.
Typical Fillers for Polymers 23

Thermo-oxidative process : lamp black

Refractory bricks

Smoke treatment for


carbon black recovery

Cooling water
Flame under regulated air
admission (=>partial
combustion of feedstock)
Air Air
Oil (feedstock)

Figure 4.2
Carbon black manufacturing process for lamp black.

After filtration, carbon black aggregates are first packed into agglomerates
then into pellets (of millimeter dimensions) in order to produce roughly
spherical grains that are easy to handle. The process has several advantages:
first it is a totally sealed, so that full respect of environment is obtained sec-
ond a precise control of elementary particle size (from 10 to 100 nm) and of
aggregate structures is achievable.1
The lamp black process is likely the oldest industrial process and has
consequently been the object of numerous engineering variants. Figure 4.2
describes the principle of a typical modern plant. The partial combustion of
a feedstock (oil generally) in an atmosphere purposely poor in oxygen pro-
duces smoke, which is cooled down and filtered to recover carbon black par-
ticles that are subsequently flocculated. The control of the pyrolytic process
is loose and results in a large distribution of elementary particle sizes (from
60 to 200 nm). This fabrication process tends to be abandoned today in favor
of the much cleaner and more versatile furnace one.
Degussa (now EVONIK), in Germany, developed the so-called gas black
process in the 1930s. Initially coal tar was used, quite a common feedstock
at that time, when carbochemistry was very important in a country that
had limited access to petrol. Today, any kind of hydrocarbon feedstock
may be used in the process. As shown in Figure 4.3, a carrying gas is flown
over preheated oil, then the oil–rich gas feeds a burner. Smoke is in part
captured on the wall of water-cooled rotating cylinders and removed with
scrapers, and in part recovered through filtration. Very fine elementary
24 Filled Polymers

Thermo-oxidative process : gas black

Off gas

Cooling water Rotating drum

Knife
Carrier gas Air

Oil
(feedstock)
Heating gas Burner

Carbon black

Figure 4.3
Carbon black manufacturing process for gas black.

particles are obtained, in the 10–30 nm range, which aggregate in a control-


lable manner.
The thermal black process is discontinuous and consists essentially of
“cooking” a mixture of natural (i.e., hydrocarbons) and inert (N2) gas in two
reactors in tandem, where cycles of heating then decomposition are achieved
(Figure 4.4). One of the reactors is heated for five to eight minutes by burning
natural gas in presence of air, whilst the other, previously heated, is loaded
with pure natural gas that thermally decomposes. When pyrolysis is com-
plete, flushing the reactor and conveying the smoke to filtration equipment
allows the recovery of carbon black particles. Depending on the ratio natu-
ral/inert gas, various ranges of large and gross particles are obtained, either
from 120 to 200 nm or from 300 to 500 nm.
Acetylene black is produced by pyrolysing acetylene at high temperature;
essentially hydrogen and carbon are obtained. Very pure carbon black par-
ticles are obtained in the 30–40 nm range.
All the processes described above yield a very wide range of carbon blacks,
differing in a number of properties, as described in Table 4.3.

4.1.3  Structural Aspects and Characterization of Carbon Blacks


In terms of consumption, the most significant usage of carbon blacks is
rubber reinforcement. This effect was discovered in the early years of the
twentieth century and played an essential role in the development of tire
technology and consequently in the automotive industry. As we shall see
later, there are still some unknown aspects of carbon black reinforcement
Typical Fillers for Polymers 25

Thermal process
H2

Gas Air

Chimney

Carbon
Ovens black

Natural gas

Heating cycle Decomposition cycle

Figure 4.4
Carbon black manufacturing process for thermal black.

Table 4.3
Carbon Black Production: Properties vs. Process
Thermo-Oxidative Thermal
Decomposition Decomposition
Lamp Gas Furnace Thermal Acetylene
Property Unit Black Black Black Black Black
Specific area, N2 m2/g 16–24 90–500 15–450 6–15 Around 65
Adsorption, I2 mg/g 23–33 n.a. 14–50 6–15 Ar. 100
Particle size nm 110–120 10–30 10–80 120–500 32–42
DBP absorption ml/100g 100–120 n.a. 40–200 37–43 150–200
Oil absorption g/100g 250–400 220–1100 200–500 65–90 400–500
Volatiles % 1–2.5 4–24 0.6–6 0.5–10 0.5–2
pH 6–9 4–6 6–10 7–9 5–8

but a basic consideration is surely the capability of carbon black to exhibit


various levels of spatial organization, which are schematically outlined in
Figure 4.5.
Elementary particles of approximately spherical shape appear in the early
stages of the pyrolysis process, when soot is being formed and assembled
together through partial fusion to give complex tri-dimensional objects,
called aggregates. During quenching, aggregates entangle into agglomer-
ates, which are eventually pelletized into granules of millimeter dimensions.
Aggregates are very difficult, if not impossible (in rubber mixing conditions)
to break and consequently are likely to be the ultimate particle size when
26 Filled Polymers

Elementary particle Occuring in soot


(colloidal black) Partial fusion formation and Carbon blacks
quenching
Occuring during
10–90 nm Entanglement wet filtering and
Aggregate
powder drying

Agglomerate Compaction
100–300 nm

Pellet
Specific
area 2–4 mm
104 –106 nm

“Structure”

Reinforcing character

10–8m 10–7m Scaling 10–5m 10–3m

Figure 4.5
Spatial organization and reinforcing character of carbon blacks.

dispersion is optimal. They play an essential role in rubber reinforcement


whilst residual agglomerates are considered as failure initiation sites in filled
compounds. There is a consensus nowadays to consider aggregates as the
ultimate carbon black form, the only relevant one when rubber (polymer)
reinforcement is concerned.
Carbon blacks exhibit various characteristics whose importance depends
on the application considered. The specific area (m2/g) is obviously the most
basic information for a given filler: the smaller the elementary particles, the
higher the specific area of aggregates with equivalent spherical volume. The
specific gravity of carbon black is in the 1.82–1.89 g/cm3 range. Note that the
common value of 1.86 g/cm3 will be used in all the illustrative calculations
made in this book.
The elementary analysis of carbon blacks is roughly as follows (in wt%):

Carbon: 95.0–99.5
Hydrogen: 0.2–1.3
Oxygen: 0.2–0.5
Nitrogen: 0.0–0.7
Sulfur: 0.1–1.0
Ashes: below 1.0
Typical Fillers for Polymers 27

In addition, toluene extraction reveals traces of organic materials, essen-


tially poly-aromatic hydrocarbons (below 0.5 wt%). A number of oxygenated
chemical groups have been found on carbon black surface, such as carbonyls,
carboxyls, pyrones, phenols, quinone, lactol, etc., but in minute quantities
and all are removed by heating at 950°C in an oxygen free atmosphere.
A number of standard characterization methods for carbon blacks (and
other fillers) are listed in Table 4.4; most of them are described as International
Organization for Standardization (ISO), American Society for Testing and
Materials (ASTM), or Deutsches Institute für Normung (German Institute for
Standards) (DIN) methods. One can distinguish three groups of methods with
respect to the information sought for reinforcement purposes: specific area,
structure and chemical analysis. In addition, there are methods for character-
izing the final product, i.e., the carbon black in pellet form, the only one read-
ily handled in polymer technology (essentially for health and safety reasons).
The specific surface area is assessed either through iodine I2 adsorption
(result is given in mg of I2 per g of carbon black), or through nitrogen N2
adsorption (result in m2/g of carbon black), or through cetyltrimethylammo-
nium bromide CTAB adsorption (result in m2/g of carbon black) or through
the tint strength (an indirect measure of the specific area). As expected, all

Table 4.4
Standard Characterization Methods for Fillers
Method ISO ASTM DIN
Specific area
Iodine I2 adsorption 1304 D-1510 53582
Nitrogen N2 adsorption 4652 D-3037/4820 66132
CTAB adsorption 6810 D-3765
Tint strength 5435 D-3265

Structure
DBP absorption 4656 D-2414 53601
Compressed DBP absorption 6894 D-3493 787/5
Oil absorption* 787/5

Chemical analysis
Volatiles 53552
Ashes 1125 D-1506 53586
Moisture* 787/2 D-1509 787/2
Sieving residue* 787/18 D-1514 787/18
Toluene extraction 53553
pH* 787/9 D-1512 787/9

Final product properties


Bulk specific gravity 1306 D-1513 53600
Pellet hardness D-3313
Pellet sizes distribution D-1511
* DIN-ISO methods.
28 Filled Polymers

of these methods give similar but not equivalent information and there is
no consensus regarding their respective advantages/disadvantages. Iodine
adsorption for instance is sensitive to surface chemistry and the presence
of polyaromatic hydrocarbons, but the method is considered correct for fur-
nace and lamp blacks. The N2 adsorption is the BET method whose principle
is based on the shape of adsorption isotherms.2 When a monolayer of mate-
rial is adsorbed on a very uniform surface, a knee occurs in the isotherm
before reaching another plateau that corresponds to the adsorption of a sec-
ond layer. Ordinary surfaces are energetically quite heterogeneous as far as
the adsorption energy in the first layer is concerned: however it is possible to
work with carbon blacks because particles exhibit (at least locally) graphite-
like facets that are quite homogeneous. Nevertheless the nitrogen molecule
is small enough to reach pores and other small cavities within aggregates,
and therefore results are sometimes obtained in excess with the area readily
in contact with the polymer matrix. CTAB is a larger molecule than nitro-
gen and consequently only the “external” specific area is probed, leading to
results better correlated with aggregates’ size.
All methods for structure assessment are indirect and essentially consist
of measuring the absorbed amount of a suitable chemical, for instance dibu-
tylphthalate DBP. Results are expressed in ml or cm3 (of DBP) absorbed per
100 g of filler. The method consists of adding dropwise DBP to a known
quantity of carbon black, which is malaxed in a calibrated laboratory mixer.
Mixing torque is recorded and as long as the liquid just fills the “voids”
between aggregates, the torque trace remains essentially flat. As soon as the
whole external surface of aggregates is “wetted,” a coherent mass is obtained
and a significant torque rise is observed. The more complex the structure,
the higher the amount of absorbed DBP. As such, the method does not dis-
tinguish between aggregates and agglomerates, and this limitation is some-
how overcome with the so-called compressed (or crushed) DPA absorption
method. Before loading the mixer cavity the carbon black is compressed four
times under a pressure of 24 MPa. Only the permanent structure, i.e., the
aggregates, is expected to survive the crushing step.
The size of the elementary particles and the structure of aggregates are
the most important parameters in the ability of a given carbon black to rein-
force a polymer. And both parameters essentially depend on the fabrica-
tion process. In time, a number of manufacturers went on the market with
products, essentially described with respect to their manufacturing process
and (expected) reinforcing character, essentially with respect to tire applica-
tions. For years, carbon blacks were described through acronyms, such as
HAF, i.e., high abrasion furnace: a furnace black imparting a high resistance
to abrasion to rubber (tread band) compounds, or ISAF-LS, i.e., intermediate
superabrasion furnace-low structure: a furnace black of low structure offer-
ing an excellent resistance to abrasion, to name a few. Only carbon black
experts familiar with rubber reinforcement aspects—and aware of the
(sometimes) subtle differences between carbon blacks, described by similar
Typical Fillers for Polymers 29

acronyms but produced by different companies—were at ease with such a


description. The essential role played by carbon blacks in rubber reinforce-
ment prompted the American Society for Testing Materials (ASTM) to pro-
pose a standard classification and nomenclature, described in ASTM D1765
that became widely accepted, thanks to its simplicity. Figure 4.6 illustrates
the principle of this nomenclature that invites all carbon black manufactur-
ers to class their materials using a four character system: one letter (either
N or S) and three numbers. The letter N means normal curing, meaning
that the carbon black does not interfere (too much) with vulcanization
chemistry; S means slow curing and concerns carbon blacks prepared with
feedstock leaving chemical residues that affect the vulcanization process. S
grade carbon blacks tend to disappear nowadays. The first number refers to
the size of the elementary particles, at least in the 1986 version of the stan-
dard, because a recent proposal (in 1996) was made to assign the first digit
a value (between 0 and 9) with respect to the specific area, as measured
through N2 adsorption. The two last digits refer to the aggregate structure
and are assigned to the carbon black by the manufacturer, with respect to
various evaluation techniques, presently not standardized however.
As a matter of fact, a very large diversity of carbon black grades has been
and is still produced, essentially because there is an excessively large num-
ber of process variables, not all perfectly monitored despite encouraging
progresses, as well as a great variety in the feedstock used. Despite modern

Refers to aggregate structure

N x yz

Refers to the size of elementary particles


Change in
defining x D1765–86 D1765–96

Typical average size Average nitrogen N2


x of particles, nm specific area, m2/g
0 1–10 > 150
1 11–19 121–150
2 20–25 100–120
3 26–30 70–99
4 31–39 50–69
5 40–48 40–49
6 49–60 33–39
7 61–100 21–32
8 101–200 11–20
9 201–500 0–10

Figure 4.6
ASTM classification of carbon blacks.
30 Filled Polymers

trends in standardization, this inevitably results in a large diversity of prod-


ucts. The present ASTM classification schema obviously offers a number
of advantages but only the first digit (X in Figure 4.6) can be considered as
solid information; in other words, for a given carbon black grade, its ASTM
classification guarantees only the size range of the elementary particles.
The structure that, in the opinion of the author, is likely the most significant
aspect is “described” by the two last digits (yz). This description is however
depending on the set of methods used by the manufacturer. To document
this aspect, Table 4.5 was filled by compiling and averaging out typical
test data for carbon black, as found in the trade or scientific literature, with
respect to their ASTM designation. Sources of data, as well as a few interest-
ing relationships between the quoted quantities are given in Appendix 4.1.
Practical experience, essentially with respect to applications in tire tech-
nology, allows to somewhat distribute available carbon blacks in three main
categories:

• Highly reinforcing, so called “tread” blacks: series N100–N300


• Semireinforcing, so called “carcass” blacks: series N300–N600
• Weakly reinforcing: series N600, N700

The words “tread” and “carcass” refer to tire applications and with respect
to wearing resistance, it is clear that tread band compounds need highly
reinforcing blacks. As a rule of thumb, the higher the reinforcing capabili-
ties of a carbon black, the more difficult its dispersion in the rubber, and
consequently the more complex the mixing procedure. Conversely, low rein-
forcing blacks can be added to rubber formulations in very large quantities.
An attractive manner to consider carbon black grades consists of plotting a
parameter related to elementary particle size, for instance CTAB adsorption,
vs. a parameter related to aggregate structure, for instance DBP absorption
(Figure 4.7). Roughly speaking the reinforcing character increases as one
moves along the increasing left to right diagonal. As can be seen in Figure 4.7,
not all combinations of both parameters are available, and the largest variety
of grades is found in the N300 series.

4.1.4  Carbon Black Aggregates as Mass Fractal Objects


The central role plaid by aggregates in the reinforcing capabilities of carbon
black is nowadays well established but recent progress in particles observa-
tion techniques, as well as fundamental studies on the physics of elemen-
tary particle aggregation through ballistic processes occurring during soot
formation and quenching, shed new light on the particular nature of such
complex objects. New concepts such as “fractal objects,” obviously inspired
by the breakthrough work of B. Mandelbrot,3 were considered in describing
carbon black aggregates.
Table 4.5
Carbon Blacks—ASTM Designation vs. Characterization Data, as Compiled from Trade and Scientific Literature
Data: DBP Absorption Comp. DBP ab. 24M4 N2 Adsorption I2 Adsorption CTAB Adsorption
Unit: (dm³/kg) (dm³/kg) (m²/g) (mg/g) (m²/g)
Method: ASTM D2414 ASTM D3493 ASTM D3037 ASTM D1510 ASTM D3765
ASTM Nr Mean value Std. dev. Mean value Std. dev. Mean value Std. dev. Mean value Std. dev. Mean value Std. dev.
       
N110 1.14 0.02 0.98 0.04 137.97 4.82 145.60 2.00 125.45 3.77
Typical Fillers for Polymers

N115 1.13 0.00 0.97 0.01 143.00 – 156.50 4.95 128.00 –


N121 1.33 0.01 1.10 0.03 126.00 5.29 121.00 0.00 118.50 3.54
N125 1.09 0.07 0.95 0.08   121.00 – 126.00 –
N134 1.27 – 1.02 – 152.00 – 142.00 – 134.00 –
N210 0.85 –        
N219 0.73 0.00   – 112.67 6.35 117.00 –  
N220 1.15 0.02 0.98 0.02 116.75 4.19 119.93 2.24 111.56 3.50
N231 0.92 0.00 0.85 0.02 117.00 0.00 121.00 – 108.50 0.71
N234 1.24 0.02 1.01 0.03 123.26 1.97 121.28 1.86 118.63 1.86
N242 1.37 0.07   – 128.75 9.91 119.00 – – –
N299 1.23 0.01 1.05 0.00 106.50 1.80 108.93 1.62 105.90 2.69
N326 0.76 0.12 0.68 0.02 82.57 8.22 83.59 2.69 82.82 4.52
N330 1.02 0.03 0.87 0.02 79.45 4.28 81.50 1.36 80.54 2.62
N339 1.20 0.04 0.99 0.04 92.10 4.32 90.15 1.22 91.31 5.91
N343 1.30 – 1.04 – – – 92.00 –
N347 1.25 0.05 0.97 0.03 87.14 3.50 89.29 2.75 85.40 3.10
(continued)
31
Table 4.5  (Continued)
32

Data: DBP Absorption Comp. DBP ab. 24M4 N2 Adsorption I2 Adsorption CTAB Adsorption
Unit: (dm³/kg) (dm³/kg) (m²/g) (mg/g) (m²/g)
Method: ASTM D2414 ASTM D3493 ASTM D3037 ASTM D1510 ASTM D3765
ASTM Nr Mean value Std. dev. Mean value Std. dev. Mean value Std. dev. Mean value Std. dev. Mean value Std. dev.
       
N351 1.21 0.02 0.96 0.02 71.43 2.17 68.22 1.09 73.23 0.52
N356 1.57 – 1.14 – 88.05 5.59 82.90 – 87.50 –
N358 1.55 0.03 1.13 0.01 84.30 2.05 84.00 – 88.00 –
N375 1.15 0.01 0.90 0.10 93.48 11.01 88.72 3.45 91.72 10.29
N472 1.86 0.08 1.18 0.06 248.67 18.48 250.00 – 147.50 3.54
N539 1.03 0.11 0.83 0.01 42.70 1.57 42.75 0.50 41.33 0.65
N550 1.17 0.14 0.84 0.04 41.12 2.82 42.58 1.11 41.32 1.21
N650 1.24 0.04 0.83 0.02 35.75 3.18 35.00 1.41 36.00 –
N660 0.94 0.12 0.73 0.03 34.85 3.09 35.73 2.02 37.52 3.04
N683 1.29 0.17 0.86 0.02 38.12 2.04 34.54 2.75 40.24 2.12
N762 0.66 0.02 0.58 0.03 28.47 3.49 28.20 1.85 29.75 5.03
N765 1.17 0.04 0.81 0.05 33.88 3.06 32.70 2.40 36.17 3.96
N770 0.73 0.04   25.00 2.00 28.00 –  
N772 0.65 0.00 0.59 0.01   30.00 0.00 33.00 –
N774 0.74 0.02 0.63 0.02 29.64 2.48 29.05 0.10 30.37 2.28
N880 0.35 0.08   12.25 2.87   –  
N990 0.38 0.04 0.37 0.02 9.09 0.93 9.40 1.22 9.70 0.99

Note: In the column for standard deviation – means that only one source of data was available.
Filled Polymers
Typical Fillers for Polymers 33

140.0

120.0 N110
N234
N220 N299
CTAB adsorption (m2/g)

100.0 r N339
racte
ha N347
80.0 N326 ng c N330 N356
ci
in for N351
Re
60.0

N550
40.0 N683
N660
N774
20.0 N762
N990
0.0
0.00 0.20 0.40 0.60 0.80 1.00 1.20 1.40 1.60
DBP absorption (dm3/kg)

Figure 4.7
Carbon blacks reinforcing capabilities with respect to parameters related to elementary par-
ticle size and aggregate structure.

Carbon black aggregates can indeed be viewed as mass fractal objects


whose description results from the so-called “fractal scaling law”: two
parts of a fractal object, a larger one of size DL and a smaller one of size DS, are
­statistically equivalent if the latter is enlarged by a factor DL/DS. Applied to the
case of an (carbon black) aggregate of overall size D made of Np aggregated
­elementary particles of size d, such a law leads to the following equality
(see Figure 4.8a):

F
 D
Np = α   (4.1)
 d

where α is a prefactor, also called front factor and F is the so-called mass-
fractal dimension of the aggregate, which depends on the conditions for the
aggregation process.
The mass fractal F describes how the mass of an object varies with its size.
This concept was first applied by Kaye4 and Flook5 to the determination
of the perimeter type fractal of carbon black aggregates, then was used by
Bourrat et al.6 Ehrburger-Dolle and Tence7 for the structural characterization
of a few carbon blacks, and by Herd et al.8 who reported an extensive study
comparing the utility of fractal and Euclidean geometries in characterizing
quite a large series of 19 carbon black grades, with DBP adsorption ranging
from 35 to 174 cm3/100 g. By measuring the perimeter and the mass fractal
values for various carbon blacks in the dry state, these authors found that
34

(a) (b) (c)

Size D Geometrical
distance R Length L

Fractal
path L
Size d

Size d
Np particles Size d

Figure 4.8
Fractal geometry description of aggregates (a) basic dimensions of an aggregate, (b) concept of fractal path, (c) Chain-like aggregate.
Filled Polymers
Typical Fillers for Polymers 35

the mass fractal dimension F was in the 2.19–2.85 range, i.e., a mean value
of 2.44±0.15. This indicates that carbon blacks have a moderately rough sur-
face, since a smooth surface has a value of F = 2. Indeed, Göritz et al.9 used
scanning tunneling microscopy (STM), atomic force microscopy (AFM) and
small-angle x-ray scattering (SAXS) to study the surface structure of quite
a broad range of carbon blacks, from N115 to N990. They well documented
the surface topography of carbon blacks and demonstrated that graphitized
(2700°C treatment) high structure blacks, e.g., N115 and N234, lost all sur-
face roughness and exhibit typical flat huge local terraces. Moreover x-ray
scattering experiments gave access to the surface fractal dimension which
was found to vary systematically from 2.27 (N115) down to about 2.0 (N990).
The smaller the primary particle diameter the higher the measured frac-
tal dimension. Consequently surface roughness decreases with increasing
primary particle diameter and is related to the reinforcing character of car-
bon black grades. One would therefore expect fractal dimensions of carbon
blacks to be correlated with DBP absorption numbers in the dry state. Figure
4.9 is for instance drawn using mass fractal dimensions reported by Herd
et al. and average DBP absorption data from Table 4.5. Mean DBPA data
and their standard deviation were used in drawing the graph. If, indeed,
there is a loose linear correlation between both characteristics of carbon
black, one can hardly expect the measurement of mass fractal dimensions to
become a valid replacement candidate for the well spread and much easier
ASTM methods, particularly with respect to the complexity of the former.

2.0

1.5
DBPA, dm3/kg

1.0

0.5

0.0
2 2.2 2.4 2.6 2.8 3
Mass fractal dimension F

Figure 4.9
Mass fractal dimension vs. (mean) DPA absorption number of carbon black. (Mass fractal data
from C.R. Herd, G.C. McDonald, R.E. Smith, W.M. Hess., Rubb. Chem. Technol., 66, 491–509, 1993.
DBPA data from Table 4.5.)
36 Filled Polymers

Table 4.6
Surface Energy Components for Carbon Black
Carbon Specific Surface Dispersive Polar Component
p
Black Area N2 (BET) Component γ ds sp
I benzene ( γs )
Grade [m²/g] (at 150°C) [mJ/m²] (at 150°C) [mJ/m²] Source
N110 140.0 270.4 120.0 29
N220 118.0 235.2 103.9 29
N234 123.3* 382.0* 93.0* 55
N326 83.2 186.5 90.2 29
N330 76.5; 80.0 196.9; 150.4 85.9; 80.2 29,30
N347 85.8 192.9 87.9 29
N550 39.7; 43.2 134.4; 173.4 75.0; 75.0 29,31
N660 39.4 124.7 71.1 29
N762 32.5 ; 24,0 126.4  ; 132.8 77.7 ; 74.0 29,31
N774 29.0 118.1 63.8 30
N880 12.3 113.1 63.9 31
N990 7.9; 10.3 71.8; 78.7 56.6; 58.8 29,31

* BET value from Table 4.5; γ ds and I benzene


sp
measured at 180°C.

Nevertheless, the fractal description of carbon black aggregates on one hand


brings quite a convincing theoretical support for the former interpretation
of DBP absorption results by Medalia and, on the other hand, provides the
starting argument for several recent theoretical descriptions of certain non-
linear effects associated with the reinforcement of rubbers by carbon black.
As largely illustrated by published transmission electron micrographs (see
Herd et al.8,10 for instance), most carbon black aggregates exhibit a branching
structure that can be considered in terms of fractal geometry. If L is the short-
est connecting (fractal) path between any two arbitrarily chosen elementary
particle of an aggregate (see Figure 4.8b), then this quantity is related to their
geometrical distance R in the three dimensional (Euclidean) space through
the following relationship:

C
L  D
= β  (4.2)
d  d

where β is a prefactor of the order or unity and C, the so-called


­connectivity exponent, readily related to the branching structure of the
aggregate. Indeed for a chain-like aggregate (i.e., without any branches;
see Figure 4.8c), the fractal path is the length of the chain and conse-
quently the connectivity exponent C equals the mass fractal dimension F
and has the value of 2. It follows that, because they have many branches,
most carbon black aggregates have connectivity exponent C significantly
Typical Fillers for Polymers 37

smaller than two. In fact, when performing their TEM/AIA (transmission-


electron-­m icroscopy/automated-image-analysis) study on a representa-
tive sampling of 19 different carbon black grades, Herd et al. 8,10 classified
aggregates in four specific shapes: spheroidal, ellipsoidal, linear, and
branched. Different aggregate shapes exist within a given grade of carbon
black but it seems that the highest percentages of branched aggregates are
found in the highly reinforcing carbon black grades. In weight percent, the
branched aggregates quantity decreases as both the DBP absorption num-
ber and the surface area decrease. With respect to reinforcement, branched
aggregates have the greatest influence on properties such as modulus, tear,
and wear resistance, which are known to be connected with the effective-
ness of the aggregate in “occluding” the polymer from deformation. The
(mass) fractal description of carbon black suits obviously the effects of
aggregate branching and, as we will see, is the background of advanced
modeling approaches.
The size D of an aggregate can be viewed as the diameter of its spherical
envelope with respect the well-known void volume concept of aggregates
introduced by Medalia.11,12 Whilst apparently not aware of the concept
of “fractals,” it is quite clear that Medalia somewhat foresaw the fractal
nature of carbon black aggregates when he wrote: “the effective volume of
a carbon black aggregate composed of Np particles is proportional to Np raised
to a power greater than unity, so that with increasing number of particles per
aggregate, the aggregate becomes more open and more voluminous.” The effec-
tive volume of an aggregate cannot however be directly assessed with
any precision and, therefore with respect to the at-the-time capabilities of
electronic microphotography techniques, Medalia suggested to consider
that the effective volume of an aggregate is that of a sphere of the same
(mean) projected area as the aggregate (see Figure 4.10). If Des is the diam-

Equivalent sphere
of diameter Des
Size D

seen by TEM

Size d Project area A


Np particles of the aggregate

Figure 4.10
Concept of equivalent sphere for a single aggregate.
38 Filled Polymers

eter of the equivalent sphere and A is the measured projected area of the
aggregate, it follows that:

3/2
π Des3 π  4 A  4 A 3/2
Ves = =  = (4.3)
6 6  π  3 π

Obviously the volume of solid carbon within an aggregate is the product


of the number of particles Np times the volume of an elementary particle
(assumed to be spherical). It follows that the measured projected area (a two
dimensional quantity) of an aggregate can be related to the project area of its
calculated equivalent sphere (a three dimensional quantity) through a scal-
ing law. Medalia et al. performed a so-called “floc simulation” to establish the
following equality:13,14

1/ε
 A
Np =  or A = Ap N pε (4.4)
 Ap 

where Np is the number of elementary particle of projected area Ap and ε a


scaling exponent. From his floc simulation, Medalia reported a value of 0.87
for ε , which however seems to be an unfortunate printing mistake since,
using his published data, it can be shown that the correct value is 0.847 (see
Appendix 4.2). If d is the (average) diameter of the elementary particles of the
aggregate, it follows from Equation 4.4:

1/ε
π d2 ε  4A
A= N p or N p =  (4.5)
4  π d 2 

Using ε = 0.847 , Equation 4.5 is rewritten as N p = ( 4 A/πd 2 )1.18 , with an


exponent slightly different from the one reported by Medalia (i.e., 1.15),
not only in his original publication but also in all his subsequent ones,
and moreover blindly used by a number of other authors. It is worth not-
ing that the exponent 1.18 is still far, but closer to the surface fractal expo-
nent of ≈ 1.8 as derived later from colloid agglomeration modeling,15,16 and
also confirmed by experimental results on carbon black filled Ethylene-
Propylene-Diene Monomer rubber (EPDM) compounds.17
From Equations 4.3 through 4.5 it follows that the diameter of the equiva-
lent sphere is given by:

Des = d N pε/2 or Des = d N p0.4235 (4.6)


Typical Fillers for Polymers 39

The exponent in Equation 4.6 is sufficiently different from the one reported
by Medalia (i.e., 0.435) to bring large differences in calculated Des when either
the diameter d of the elementary particle increases and/or when the number
Np of elementary particles is large. However, the misprint in Medalia publica-
tions must not shade his merit in having foreseen the fractal nature of carbon
black aggregates. In addition, Medalia has thoroughly elaborated practical
formulas to convert an easily measured quantity (i.e., the DBP absorption
number, in cm3/100 g filler) into the number of particles in an aggregate (see
Appendix 4.3 for details and numerical illustrations).
The solid volume Vs of an aggregate is nothing else that the volume of an
elementary particle (of diameter d) times the number of particles, and by
combining Equations 4.3 and 4.5, it follows:
1/ε
π d3  4 A  π d3
Vs = N p = (4.7)
6  π d 2  6

Using his floc simulation approach, Medalia has established the following
practical relationship between the so-called “void ratio,” i.e., the ratio of the
equivalent sphere Ves to the solid volume Vs, and the DBP absorption num-
ber, i.e.

Ves (1 + vf )
CF g − 1 = DBPA ρ 0.0115 (4.8)
Vs C

where
CF = 0.765: correction factor accounting for difference between the pro-
jected area of the equivalent sphere and the projected area of the aggregate
(around 8.5% reduction in diameter)
vf = 0.46: void fraction for randomly packed spheres
C = 1.4: correction for partial fusion of primary particles in aggregate
g = 0.94: anisometry correction factor for non-perfect alignment of aggre-
gate’s main axis with projection plan
ρ = filler specific gravity [carbon black: ρ = 1.86 g/cm3]
DPBA = di-butylphtalate absorption (cm3/100 g filler)
0.0115: correction for DBPA end point (i.e., 1.15/100 , to take into account that
at the end of the DBP absorption test, the sample contains around 15% air)
By combining the equations for the volume of the equivalent sphere
(Equation 4.3), for the projected area (Equation 4.5) and for the volume of
solid carbon in the aggregate (Equation 4.7) with Equation 4.8, one gets
immediately:

 ε 
 ε + 2 − 1 C
N p = ( 1 + DBPA ⋅ ρ ⋅ 0.0115 ) ⋅
CF ⋅ g ⋅ ( 1 + vf )
40 Filled Polymers

or

 ε 
 ε + 2 − 1
N p = ( 1 + DBPA ⋅ ρ ⋅ 0.0115 ) ⋅ 1.333 (4.9)

if one replaces the various correction factors by their values given above.
The number of elementary particle in an aggregate can consequently be
assessed from the DBPA number, the specific gravity ρ of carbon black (1.8
g/cm3) and the value assigned to ε , using:

N p = [ 1.333 ⋅ ( 1 + DBPA ⋅ ρ ⋅ 0.0115 )] 2/( 3 ε − 2 ) (4.10)

Depending on the value used for ε , the result yielded by Equation 4.10
can be very different, as shown in Figure 4.11, in fact largely underestimated
using ε = 0.87 , as published by Medalia. For instance, for a typical High
Abrasion Furnance (HAF) grade, e.g., N330 (DBPA = 102 cm3/100 g), the cor-
rect value ε = 0.847 gives 211 particles/aggregate, whilst ε = 0.87 gives 114.
The more reinforcing the carbon black, the larger the difference.
It is interesting to compare the number of particles per aggregate as cal-
culated with Equation 4.10 (Medalia’s; based on TEM analysis and “floc”

800

700
ε = 0.847

600
Number of particles

500

400

300
ε = 0.87
200 (Medalia)

100

0
0 0.5 1 1.5 2
DBPA, dm3/kg

Figure 4.11
Assessing the number of elementary particles in a carbon black aggregate; curves were cal-
culated with Equation 4.10, ρ = 1.86 g/cm3, DBPA values in the range 38 (N990) to 157 (N356)
cm3/100 g and the ε values given in the figure. (Data (◽) are from A.I. Medalia, F.A. Heckman,
Carbon, 7, 567–582, 1969.)
Typical Fillers for Polymers 41

800

Number of particles/aggregate (Medalia)


600

400

200

0
0 200 400 600
Number of particles/aggregate (Fractal approach)

Figure 4.12
Comparing carbon black aggregates as described either through the TEM data analysis by
Medalia and or the fractal approach.

simulation) with the estimation obtained using the mass fractal approach,
i.e., Equation 4.1, and reported particles and aggregates dimensions (from
Herd et al. for instance; see Appendix 4.4 for details). As shown in Figure
4.12, with respect to the equality line, an agreement is obtained only if the
front factor α in Equation 4.1 is taken equal to around 11, i.e., more than 10
times the value guessed by some authors.18
Within the spherical aggregate envelope, one can distinguish the solid Vs
and the void Vv volumes, whose relative importance is expressed in terms of
volume fraction, i.e.:

−1
 V 
Φ = 1+ v  (4.11)
 Vs 

With respect to fractal geometry, this solid volume fraction is expressed in


terms of basic dimensions of the aggregate as follows :

3− F
N d3  d
Φ= =α  (4.12)
D3  D

where α is the so-called “front factor.” This fraction is the volume occupied
by the (fractal) solid aggregate with respect to the overall volume occupied
in the three dimensional space, and is obviously related to the aggregate sur-
face accessible to polymer chains in a compound. When carbon black volume
42 Filled Polymers

fraction is large enough in a compound—in practice when the loading is


above the so-called percolation level (i.e., Φ ≈ 0.12 − 0.13 ), all aggregates are
expected to entangle (or at least to connect) and to form a secondary aggre-
gated structure in the polymer matrix.
As we have seen before, during the production process (precisely during
the quenching), carbon black aggregates flocculate into agglomerates, that
are further compacted by the final pelletizing step (see Figure 4.5). Carbon
black pellets is the easier handling form of the filler, readily used in polymer
compounding. Such pellets are very friable and no much (mixing) energy is
needed to split them into agglomerates which correspond in fact to a close
packing state of aggregates. Agglomerates have also a fractal nature with a
mass fractal dimension of the order of three and a connectivity exponent of
around one. Aggregates are recognized for decades as the filler structural
state that plays the key role in rubber reinforcement, but some authors have
recently argued that, when above a critical concentration threshold they are
well dispersed in a polymer matrix, they form a kind of tenuous second-
ary structure, which helps in understanding certain aspects of the reinforce-
ment of elastomers through a filler aggregates networking effect (Klüppel
and Heinrich18), as we will see later in detail.
Whatever is the packing state of aggregates into agglomerates, and the
compaction degree of agglomerates into pellets, there is a certain degree of
“voids” such that the solid fraction of carbon black pellets is given by:

DBPA
Φ pellet = 1 + ρ (4.13)
100

The fraction Φ pellet is found nearly equal to the volume fraction Φ for the
aggregate (Equations 4.11 and 4.12), which means of course that the front fac-
tor α is also close to one for carbon black pellets.
The mean number of aggregates within the total volume R3 of a single
agglomerate is defined as N aa = n R 3 with n the number density of the
aggregate, i.e.

Φ
n= (4.14)
N d3

The number N aa is obviously related to the degree of interpenetration of


aggregates in each other and, with respect to Equation 4.1 it follows:

N aa = α − 3/F Φ N (( 3/F ) − 1) (4.15)

In relation with the void volume in carbon black aggregates, Medalia19 intro-
duced the (debated) concept of “occluded rubber,” defined as the fraction of
Typical Fillers for Polymers 43

polymer that has penetrated the internal void space of filler aggregates and
is thus shielded from deformation, at least partially. Medalia classified voids
in a compacted carbon black in two categories: within and between aggre-
gates, and he developed two relationships that permit to assess their relative
importance from easily measured quantities, i.e.

voids volume within aggregate


Ratio
overall agglomeratee volume

 1 + 0.02139 DBPA 
=  ( 1 − Φ ) ×  − 1 (4.16)
 1.46 

voids volume between aggregate


Ratio
overall agglomeratte volume

1 + 0.02139 DBPA
 =  Φ − ( 1 − Φ ) ×  
− 1
 1.46 

where DPBA is expressed in cm3/100 g, the factor 0.02139 is the product


ρ × 0.0115 with ρ = 1.86 g/cm3 and the factor 1.46 is related with the Medalia’s
assumption11 that, at the endpoint of an oil absorption test, the remaining void
space between aggregates is 31.5% so that 1 + ( 31.5/100 − 31.5 ) = 1.45985.
Such considerations about the fractal nature of carbon black are the back-
ground for recent theoretical developments on the very origin of the rein-
forcing effect of the filler and of certain aspects of the mechanical properties
of rubber parts. During efficient mixing operations, carbon black pellets are
expected to completely disappear and agglomerates to fully separate into
their constitutive aggregates, the latter being ultimately evenly distributed
in the rubber matrix. A “well dispersed” state can of course be considered
in terms of an even statistical distribution in a given rubber volume, but in
the opinion of the author, it is more interesting to define the ideal well dis-
persed state as the situation where all the reinforcing entities of the filler, in
the occurrence the aggregates, have developed their maximum interaction
potential with the rubber matrix. In other words, in the optimum disper-
sion state, the maximum available specific area of the aggregate is in contact
with elastomer chains. It is now easy to understand that there will be a tre-
mendous difference in carbon black effects on mechanical (and rheological)
properties of rubber compounds, depending one is in the low concentration
regime or above a critical concentration level. Below this critical concentra-
tion level, well mixed aggregates are sufficiently separated from each other
and it is essentially the surrounding rubber matrix that support and transmit
the stress. Above a critical filler level, obviously not much depending on the
grade of carbon black, there are enough aggregates for a secondary carbon
44 Filled Polymers

black network to be formed through direct aggregate—aggregate interac-


tions, with the resulting capability to support and transmit stress in the com-
pound. As local aggregate density in the rubber matrix increases, a kind of
aggregates flocculation occurs which, amongst other effects, can be consid-
ered as the very origin of phenomena such as dynamic stress softening.

4.1.5  Surface Energy Aspects of Carbon Black


In addition to specific surface area and the fractal nature of carbon black as
discussed above, it may be expected that rubber–filler interactions, which are
the roots of reinforcement, somewhat depend upon the surface activity of the
particles. The so-called surface activity is not however a clearly defined concept
as many phenomena might be involved, from Van der Waals proximity forces
(around 4 kJ/mole) to specific chemical interactions (e.g., hydrogen bond-
ing, ≈ 20 kJ/mole; ionic bonds, ≈ 30 kJ/mole). Despite the considerable literature
on the subject, there is so far no standard method to measure surface activity.
Rubber grade carbon blacks contain small quantities of chemically com-
bined hydrogen (0.2–1.0%), oxygen (0.1–4.0 %) and even sulfur (up to 1.0%)
depending on the quality of the feedstock and the process. Over the years, a
large variety of oxygen containing functional groups, most in minute quanti-
ties, has been detected in carbon blacks, for instance carboxyl and hydroxyl
groups, phenol, lactones, quinones, ketones, aldehydes, hydroperoxydes, etc.,
(Figure 4.13). It must be noted however that a number of reported data that
support the picture offered in Figure 4.13 have been obtained on lamp or gas
blacks, obviously very sensitive to contamination by oxygen and other het-
eroatoms. Advanced and sophisticated analytical techniques performed on
(modern) furnace blacks give quite a different picture. Indeed, Bertrand and

HO O
Ketone
C Carboxyl
O
O
Pyrone

O C
O
Lactone

HO
Hydroxyl
O O
Quinone

Figure 4.13
Chemical functions detected on (lamp and gas) carbon black surface.
Typical Fillers for Polymers 45

Weng20 used time-of-flight secondary ion mass spectrometry (ToF-SIMS) and


x-ray photoelectron spectroscopy (XPS) to characterize various furnace blacks,
either commercial or experimental, quite representative of the reinforcing
carbon blacks available today. They carefully interpreted the various spectra
obtained before and after toluene extraction of carbon blacks and came to the
conclusion that there are only C and H on carbon black surface, with nearly no
oxygen. Even after heat temperature treatment (1000°C), hydrogen contain-
ing fragments were still detected in ToF-SIMS spectra. This strongly supports
the view that carbon black surface is locally graphitic in nature, with broken
graphitic plan edges supporting only C–H or maybe some pendant methyl
groups. Carbon black surface chemistry therefore plays nearly no role in the
exceptional reinforcing capabilities of this filler. It is moreover well estab-
lished today that oxygen complexes at the surface of carbon black particles
are not essential for reinforcement in most rubbers, with the notable excep-
tion of polar elastomers, e.g., butyl rubber. As indeed convincingly demon-
strated by Gessler et al.21 some 30 years ago, oxygen functionality on carbon
black is a requisite to high-order interactions only for butyl rubber. Indeed
when the surface oxygen on channel black is removed by high temperature
treatment under inert gas, these authors observed significant loss in rein-
forcement (damping properties) of butyl rubber. Conversely when furnace
blacks (for butyl rubber) are activated by heating at 250–300°C in a stream
of oxygen, a significant benefit is observed in terms of reinforcement. With
nonpolar elastomers, i.e., most general purpose rubbers, including Natural
Rubber (NR), Butadiene Rubber (BR), Styrene-Butadience Rubber (SBR),
Ethylene-Propylene Rubber (EPR) and EPDM, the occurrence of chemical
reactions with functional groups on carbon black surface is far less convinc-
ing. The heating of carbon black below 800°C does not result in graphitiza-
tion of the filler, i.e., there are no significant changes in the crystallinity of the
inner particles, but it does remove most of the chemisorbed surface oxygen.
However, rubber–filler interactions remain essentially unaffected, as shown
by Dannenberg’s experiments with SBR compounds.22,23 There is therefore a
consensus today to consider that the strong rubber–carbon black interaction
is not necessarily resulting at all from chemical reactions involving oxygen
complexes at the surface of particles. However if the rubber is containing
specific reactive groups, then quite logically it may be interesting to consider
purposely surface oxidized carbon blacks to promote chemical interaction, as
shown for instance by Manna et al.24 with epoxidized natural rubber (ENR).
Indeed, with respect to a 60 phr (part per one hundred rubber) Intermediate
Super Abrasion Furnace Carbon black (ISAF) (likely N220) filled ENR refer-
ence compound, a corresponding 60 phr oxidized ISAF/ENR compound with
4 phr silane coupling agent exhibits twice higher tensile and tear properties,
largely below however with what can be readily obtained with conventional
natural rubber and carbon blacks.
It is thus well established today that carbon black surface chemistry plays
a very minor role, is any, in the reinforcement of general purpose elastomers,
46 Filled Polymers

i.e., essentially diene rubbers (NR, BR, SBR) and EPDM, more than 90% of
the overall rubber consumption. As we will see, the situation is completely
different with other fillers, namely silica, for which the surface chemistry of
particles plays the essential role.
An attempt to quantify the role of filler surface chemistry is to consider the
so-called “surface activity,” generally assessed through the surface energy γ s ,
which consists of two main components,25 i.e.

γ s = γ ds + γ sp (4.17)

where γ s and γ s are respectively the so-called dispersive and polar (or spe-
d p

cific) components.
Such properties are measured by inverse gas chromatography (IGC),26 a
technique in which the filler to be characterized is used as the stationary
phase and the injected solute is the so-called probe. In practice, the filler par-
ticles are carefully poured into a stainless steel column of appropriate diam-
eter, typically a few mm. Suitable model chemicals, in dilute solution, are
used as probes in order to quantify their interactions with the filler. When
the probe is operated at infinite dilution, information is obtained concerning
the adsorption of the solute on a solid surface, by use of the Henry’s law that
considers the standard free energy of transferring one mole of vapor from a
gas phase (at the standard pressure of 1 atmosphere, or 101 kPa) to a standard
state on the surface.27 By injecting a series of homologous n-alkanes (e.g., pen-
tane to decane) as probes, the dispersive component of the surface energy γ ds
is obtained from the free energy of adsorption, by considering the slope of
the measured standard free energy for adsorption vs. the number of carbon
atoms for different n-alkanes. The specific (or polar) component is derived
from the difference in the free energy of adsorption between a polar probe
and a real or hypothetical n-alkane with the same surface area (see details
elsewhere25,26,28–31). Table 4.6 gives typical data as reported in literature; as
usual data for the same grade slightly differ between authors.
Carbon blacks exhibit a high dispersive component, actually proportional
to their specific surface area, and a relatively low polar component, not much
differing whatever the grade. As we will see later, the reverse is seen with rein-
forcing silica grades, which have a lower dispersive component, but a high γ s .
p

With carbon blacks, the reinforcing effect is thus essentially achieved by means
of strong filler–rubber interactions, and the polar component is reflected by a
relatively weak carbon black network, at least providing that the filler volume
fraction is below the so-called percolation level. Above that level, some authors
have recently argued that branched aggregates readily entangle through com-
plex topological interactions in such a manner that the carbon black network
plays the key role in modulus enhancement, as we will see later in detail.
New equilibrium gas adsorption techniques were recently used to analyze
the surface energy distribution of carbon blacks.32–35 By deconvoluting the
Typical Fillers for Polymers 47

energy distribution function of adsorbed ethylene into four Gaussian peaks,


it can be concluded that there are four different energetic sites on the surface
of the filler (Figure 4.14). The nature of these energetic sites is however a mat-
ter of interpretation that depends on the model considered for the surface
of the filler. For instance, if one considers as with some authors36,37 that the
surface of elementary carbon black particles consists of turbostatic graphitic
crystallites beside areas of amorphous carbon, then one could assign such
energetic sites to different features:
Site I: graphitic surface of a crystallite
Site II: amorphous carbon zone.
Site III: crystallites edges
Site IV: slit shaped cavities (or boundaries between two crystallites)
Crystallite
edges Slit shaped
Amorphous
(III) cavities
carbon zones
Graphitic (IV)
(II)
planes
(I)

0.20
Adsorption of C2H2 on at T = 233 K
Distribution of surface energy f(Q), kJ/mol

N220 carbon black

0.16
I Overall energy
distribution function
0.12
Fraction (%) of energetic sites
I II III IV
0.08 N115 69 13 15 3
N220 84 7 7 2
II N550 93 6 1 <1
0.04
III Graphite 94 0 4 2
IV
0.00
0 10 20 30 40
Surface energy Q, kJ/mol

Figure 4.14
Types of energetic sites on the surface of carbon black with supporting data (From: A. Schröder.
Charakterisierung verschiedener Rußtypen durch systematische statische Gasadsorption;
Energetische Heterogenität und Fraktalität der Partikeloberfläche. PhD thesis, University of
Hannover, Germany, 2000.)
48 Filled Polymers

The fraction of higher energetic sites (ie. II and III) is then found to decrease
with the reinforcing capabilities of the black and to disappear almost com-
pletely through graphitization. It can therefore be hypothesized that these
sites play an important role in filler–filler and filler–elastomer interac-
tion. However, because of the relatively low energies of these sites, weak
interaction forces (London, Van der Waals, etc.) between carbon black sur-
face and unsaturated segments in the rubber backbone may prevail, with
obviously a significant cumulative effect due to the large number of such
sites.
With respect with the macroscopic effects involved in reinforcement,
the exact nature of the surface of carbon black particles might appear as
detail knowledge. But the key information is the demonstration of the sur-
face roughness of elementary carbon black particles, whether one considers
a disordered array of graphitic crystallites or a step like surface resulting
from a spiraling growth process. It is quite remarkable that nearly all carbon
blacks have a unique surface roughness on atomic scales with a surface frac-
tal dimension d f ≈ 2.6 , in agreement with the most recent theoretical mod-
els for physical concepts of surface growth that are nowadays found to be
valid in many different fields in nature. One such theoretical approach is the
so-called Toom model, i.e., a relatively simple probabilistic cellular automa-
ton that appears to be “generic” for a variety of physical patterns.38 In the
context of this book, such fundamental theoretical works might seems quite
remote, but it is striking to see that through calculation with the appropriate
equations, step-like surface patterns are built that are very similar to what
is observed on AFM images of carbon black.39 It seems thus that there is a
convergence of experimental and theoretical results that describe the rough
surface of elementary carbon black particles as a pattern of overlapping
scales, further reinforced by the recent demonstration of the role of fuller-
enes, particularly C60, as precursors for the chemical reactions involved in
the formation and nucleation of carbon black in the furnace combustion
process.40,41
Another indirect demonstration of the importance of the very surface of
filler particles in the development of strong interactions with polymer mate-
rials is provided by the fact that, when chemically grafting short hydro-
carbon chains on carbon blacks, the reinforcing capabilities of the filler are
either lost or at least severely penalized. Vidal et al.42 showed, for instance,
that the esterification (via carboxyl groups at the surface) of carbon blacks,
using either methanol or hexadecanol, is associated with a strong decrease
of the interactions the filler can develop with any molecule. The surface
free energy is found to decrease whilst the surface energy homogeneity is
increasing. As expected, when grafting filler particles, interparticle interac-
tions are lower, an effect that might be beneficial for dispersion mechanisms;
but polymer–filler interactions are also decreasing, which results in a severe
drop in reinforcement.
Typical Fillers for Polymers 49

4.2  White Fillers


4.2.1 A Few Typical White Fillers
So-called white fillers are materials of various nature and origin that are used
in both elastomers and thermoplastic polymers. Most of them are mere natural
products, i.e., found as such in nature and submitted to more or less compli-
cated mechanical and physical processes, such as sorting, washing, grind-
ing, sieving, etc. Others are elaborated natural products, i.e., slightly modified
through relatively simple chemical treatments, and a few are essentially syn-
thetic materials, i.e., not found as such in nature but obtained either as side
products of chemical processes or through purposely chemical synthesis.
Typical natural fillers are extracted from lodes (in quarries or mines), sorted
and washed if necessary, nearly always grinded to the appropriate particle
size (at least below 10 µm) and sieved with respect to the desired distribu-
tion of dimensions. Chemically, mineral white fillers are silicates, carbonates
or oxides, with specific surface area ranging from 1 to 50 m2/g, therefore
considered as inert (i.e., not reinforcing) to semiactive (slightly reinforcing)
materials. We briefly describe hereafter the most common white minerals
used as fillers for polymer materials.

4.2.1.1  Silicates
Kaolinites (clays): general chemical formula: Al2O3.2SiO2.2H2O. Specific
gravity: 2.6 g/cm3. Clays belong to the phyllosilicate class (old greek: ϕυλλον,
plate), which means that, at microscopic level, they exhibit lamellar forms
that are reduced to roughly hexagonal small plates by grinding (typical
aspect ratio: 20:1). Depending on the mineralogical quality of the deposit, its
geological environment and its accessibility, the extraction and treatment of
clay can be more or less complicated but the main steps of the process are
always the same: extraction (mining, excavation or high pressure water dis-
integration), grinding, separation from foreign minerals, e.g., sand, mica, etc.,
washing, sorting and screening, then thickening, filtering, drying and final
grinding. One distinguishes the “hard clays” (30 m2/g specific surface area),
which are semiactive fillers, and the “soft clays” (10 m2/g) considered as inert
fillers. The product form is a fine powder with a moisture content of around
1%. Mohs hardness is low, typically 2.5, and pH is 5–7.5 (Note that detailed
considerations on the Mohs hardness are given in Section 6.2.1).
China clay or Kaolin (from the chinese “kao-ling,” hill) is a hydrated alu-
minium silicate crystalline mineral (kaolinite) formed over many millions of
years by the hydrothermal decomposition of granite rocks. Hydrous kaolin
is characterized by its fine particle size, plate like or lamellar particle shape
and chemical inertness. There are many production sites in China and in
50 Filled Polymers

Europe, the most important deposit is in Cornwall, England (discovered in


1745 by W. Cookworthy and initially used to produce white chinaware).
Montmorillonite is a very soft type of clay, named after Montmorillon, Vienne,
France where the first deposit was discovered in 1847, but is found in many
places worldwide. It is a hydrated Na, Ca, Al, Mg silicate hydroxide (Na,Ca)0,33
(Al,Mg)2 (Si4O10)(OH)2.nH20 whose particles are plate-shaped (average diam-
eter: 1 µm) and made of silicate sheets spaced by 1–2 nm. Montmorillonite has
many usages, for example as component of drilling mud, as soil additive (due
to its high capability to absorb water), and as animal feedstuff component
(anticaking agent) and became recently a popular component in many recent
research works on so-called nanocomposites, thank to its capabilities to be
relatively easily exfoliated by polymer chains. Exfoliated montmorillonite
provides to certain polymer compositions interesting properties at very low
loading, in the 2–5% volume fraction range, which however have not yet led
to sizeable industrial applications, to the author’s knowledge.
In connection with the on-going research efforts on so-called polymer
nanocomposites, the surface modification of clay is worth to be briefly men-
tioned. Recently Liu43 published a comprehensive review on the advances in
surface modification of natural clay minerals with polymers, including the
modifying methods and mechanisms, characterizing and analytical tech-
niques, and potential applications. Two main approaches are so far exploited:
either physical adsorption or chemical grafting of functional polymers on
the surfaces of the clay minerals. Physical adsorption, a process essentially
driven by thermodynamic effects can somewhat modify the nature of the
mineral surface and improve the local physical and chemical properties,
without modification of the structure of the clay. However the forces between
the adsorbed molecules and the clay mineral might be weak, and the pro-
cess is essentially reversible, for instance nearly annihilated as temperature
increases. Coated clays, with fatty acids or oleates for instance, belong to this
physical adsorption approach and are readily used in the current industrial
practice of filled polymers. Grafting of functional polymers to the surface
of clay minerals induces stronger interactions, in principle irreversible and
therefore less sensitive to temperature. Two strategies can be used: either
one-step or two-steps grafting methods (Figure 4.15). The former consists
in the condensation of functionalized polymers with reactive groups of clay
particles. Relatively loose polymer brushes are obtained because chemisorp-
tion of the first fraction of chains hinders the diffusion of the following
chains to the surface for further attachments. The two-step grafting tech-
nique allows higher grafted densities to be achieved, because a monolayer
of polymerizable (macromonomer) or initiator (macroinitiator) molecules
is first covalently attached to the clay surface, then further activation pro-
duces chains growth from the interface, which is limited by the diffusion
of monomers to the active species only. Chemical grafting of clay minerals
to produce so-called organo-clays are complicated and expensive processes
however and there is no doubt that they offer a route to new materials with
Typical Fillers for Polymers 51

One-step grafting

X
X + Y X
X

Two-step grafting

X Z - A - A* A Z-A
X + Y - A - A* Z - A - A* Z-A
X Z - A - A* Z-A

Figure 4.15
Mechanisms for one-step and two-step grafting reactions on mineral surfaces.

predefined structure and performance that ultimately will allow to prepare


truly advanced materials, e.g., organic/inorganic polymer nanocomposites
or even biomimetic materials. There are not yet indication for significant use
of such polymer-modified clay in material technology.
Calcined kaolins (calcined clays): obtained by treating kaolinites at temper-
ature higher than 1000°C in order to remove water and modify the structure.
An anhydrous aluminium silicate is obtained, with increased whiteness and
hardness, better electrical properties, and smoother size and shape of particles.
Calcination occurs in three steps: at 700°C the dehydroxylation of the kaolin
is complete forming a poorly crystalline metakaolin, at 980°C an amorphous
defect spinel is formed which undergoes a recrystallisation in an amorphous
glass at temperatures above 1100°C. Products obtained within this tempera-
ture range provide enhanced mechanical properties and ­chemical resistance
in rubber compounds, especially when coated with silane. Calcined kaolin
improves the thermal properties of agricultural films, gives better electrical
properties in PVC cables and, suitably coated, may act as a functional filler
in engineering thermoplastics. Silane coated grades provide higher stiffness,
toughness and dimensional stability to polyamide moldings.
Talc: hydrated magnesium silicate of general chemical formula: Mg8(OH)4.
Si8O20 or 3MgO.SiO2.H2O. Specific gravity: 2.7–2.8 g/cm3. Another phyllosili-
cate with a lamellar structure and a very low hardness. Depending on the
deposit, other minerals can be present, e.g., chlorite, dolomite, sometimes in
the few percentage range. Particle sizes range from 2 (semiactive) to 15 µm
(inert filler). Talc is a common filler for polypropylene with loading up to the
30% weight and is also used in the rubber industry as antitack agent for stor-
ing uncured rubber strips before further processing. The lamellar surface of
talc is organophilic, which means that such a mineral has a natural affinity
for organic substances, including of course most hydrocarbon polymers.
Mica: a group of phyllosilicates exhibiting near perfect basal cleavage,
owing to the hexagonal sheet-like arrangement of its atoms. Muscovite is
potassium aluminum silicate hydroxide fluoride, KAl2(AlSi3O10)(F, OH)2, and
52 Filled Polymers

the most common of the mica group. It has a layered structure of aluminum
silicate sheets weakly bonded together by layers of potassium ions, which
allow a perfect cleavage in thin sheets of flakes, which are flexible and elas-
tic. Color is white to colorless, with a vitreous to pearly luster. Mohs hard-
ness is 2–2.5 and specific gravity is 2.75–2.90 g/cm3. Properly delaminated,
milled and sorted, mica from muscovite is used in certain thermoplastics as
a reinforcing filler. Biotite, another common phyllosilicate within the mica
group, with the formula K(Mg, Fe)3AlSi3O10(F, OH)2, is less used as a filler for
polymers, owing to its dark brown color. Because it is frequently associated
with other minerals, scrap and flake mica is produced all over the world but
the largest deposits are in India, China, and Brazil.

4.2.1.2  Natural Silica


Quartz: nearly pure SiO2 is obtained by finely grinding quartz, the most
abundant mineral on earth. Specific surface area (BET) is below 5 m2/g,
which confines such a silica in the class of essentially inert fillers. The high
hardness (7 on Mohs scale) severely limits its use as filler for thermoplas-
tics owing to wearing problems on processing equipment. However, certain
grades find uses in specific elastomer applications.
Neuburg silica: a natural mixture of lamellar kaolinite and corpuscular
quartz, that forms a loose structure. Specific gravity: 2.6 g/cm3. It is mined
in the region of Neuburg (Germany), and treated by hydrocyclonisation to
yield nearly spherical grains (1–5 µm diameter) that exhibit a grape structure
(BET: 11 m2/g). Mohs hardness is up to seven for the quartz particles, but
somewhat tampered by the lower harness (around 2.5) of kaolinite. Available
from Hoffmann Minerals GmbH, Neuburg, Germany, under the tradename
Sillitin®, typical grades have 65/30 to 78/17 %silica/%kaolinite ratio with
around 5% other minerals. It is a very specific filler considered as inert to
semiactive44 and is mainly used in the rubber industry.
Wollastonite: a white calcium silicate (CaSiO3), with sometimes traces of
Fe, Mg or Mn substituted for Ca, mined in certain regions (China, India, USA,
Germany, Sweden), which has a typical acicular structure (i.e., packed micro-
scopic needles), with diameter-to-length ratio ranging from 1/3, 1/5 to 1/15.
Mohs hardness between 4.5 and 5; specific gravity: 2.87–3.09 g/cm3. Wollastonite
is used as a filler for thermoplastics and is the essential white filler for fluoroelas-
tomers (FKM). With respect to its unique structure, wollastonite could also be
considered as a natural short mineral fiber. Micronized wollastonite grades are
available with very high aspect ratio (L/D), which are especially designed for
thermoplastic and thermoset applications where a high degree of reinforce-
ment is needed. Such grades could in principle be used instead of glass fibers,
providing the adequate surface treatments are made with respect to the proper-
ties of the polymer matrix. In thermoplastics compounding Polycarbonate (PA)
6 and 66, Polypropylene (PP), Polycarbonate (PC), Polyurethane (PU), wollas-
tonite is expected to provide higher heat distortion temperature, better impact
Typical Fillers for Polymers 53

and scratch resistance and less shrinkage of injection molded parts, all benefits
associated with its needle-shaped particle structure. Most wollastonite-filled
compounds are used by the automotive sector (interior and exterior parts, under-
the-hood components) and the electrical industry (insulation materials). There
are some ultrafine wollastonite grades, either untreated or surface coated modi-
fied, which no longer have anisotropy related reinforcing performance; such
grades are used in thermoplastic polyesters and plateable/paintable nylons for
improved adhesion and reflectivity. Some specially treated wollastonite prod-
ucts exhibit conductive properties so that they could replace conductive carbon
black grades or carbon fiber in thermoplastic olefins and engineering alloys.
Miscellaneous siliceous fillers: there is a wide variety of mineral deposits
that, either as such or suitably treated, may offer fine particles of interest, with
various shape, structure and/or composition. For instance, perlite is a volcanic
siliceous rock (around 75% SiO2 and 15% Al2O3), naturally occurring as hol-
low, near spherical micro-pearls, with a water content in the 3–5% range. When
shock treated at high temperature (above 600°C), perlite expands and devel-
ops an internal rigid foamed structure. Suitably milled, grinded and sorted,
such minerals exhibit low reinforcing properties, equivalent to N600 or N700
carbon blacks, but they could be useful in certain specialty elastomers, for
instance fluoroelastomers. Their hardness might however give wear problems
in polymer processing equipments. Another example is provided by so-called
diatomaceous earths which are deposits of the skeletons/shells of microani-
mals. Microscope observation reveals a great variety of delicate and perforated
structures. The typical chemical composition of diatomaceous earth is 86%
SiO2, 5% Na2O, and some magnesium and iron oxides. The natural product can
be calcined so that the particles fuse together to produce grades of controlled
average particle size. Apart their well known usage as supports for catalyst in
chemical engineering, such siliceous products could also find niche applica-
tions as either extenders of semireinforcing fillers for certain type of polymers.
Except a few technical data in trade literature, nearly no scientific information
is available (to the author’s knowledge) about the actual use of these minerals in
polymer applications, but the chemical inertia of such “exotic” minerals makes
them worth of consideration as filler materials for rubber or plastic parts with
pharmaceutical and food applications. Generally such minerals have a high
absorption potential for hydrocarbon products owing to their organophillic
character. Therefore they could also be interesting alternatives to low molecu-
lar weight polymers as used in the encapsulation of (rubber) chemicals.

4.2.1.3  Synthetic Silica


Synthetic silica (silicon dioxide) are prepared either by pyrogenation of silica
tetrachloride or by precipitation from a solution of alkaline silicates through
the action of acids or metal salts (see next section). Synthetic silica exhibit spe-
cific surface area in the 100–200 m2/g range and are therefore active fillers.
Table 4.7 gives a comparison of some physical properties and the chemical
54 Filled Polymers

Table 4.7
Properties of Precipitated Silicic Acids (Silica) vs. Active Silicates
Precipitated Silicic Acids Silicates
Physical Data Unit Very Active Active Active
Specific area (BET) m /g2 160 130 90
Moisture (2 h, 105°C) % 6 6 6
Fire loss (2 h, 1000°C) % 4 4 7
Specific gravity g/cm3 2.0 2.0 2.2
Apparent density kg/m3 250 250 250

Chemical Composition
Silicic acid SiO2.H2O % 98 90 81
Aluminium oxide Al2O3 % 0.5 8.5 8
Magnesium oxide MgO % – – 2
Sodium sulfate Na2SO4 % 1.5 0.5 –
Sodium oxide Na2O % – – 8
Sulfites SO3 % – – 1

composition of precipitated silicic acids (silica) vs. active silicates. Two major
differences emerge from this comparison: first the active character is associ-
ated with the specific surface area, second the purer the silica, the better its
reinforcing character.
The usage of silica was recently boosted in the rubber industry by the
development of the so-called “green tire” technology, pioneered by Michelin
in the late 1980s. As we will see later, high structure silica impart interest-
ing dynamic properties to tire tread compounds when compared to carbon
black, namely a lower viscous dissipation and therefore a better efficiency in
term of so-called “rolling resistance.” However a chemical treatment (using a
bifunctional silane) is necessary to achieve a good balance of properties.

4.2.1.4  Carbonates
Chalk, whiting, and limestone are natural calcium carbonate (CaCO3) that,
finely grinded, is likely the most widely available and utilized mineral as
thermoplastic additive and is sometimes used as an inert filler in rubber
compounds. Specific gravity: 2.70–2.83 g/cm3 (depending on crystal form;
calcite less dense than aragonite). Calcium carbonate ores are found in either
sedimentary or metamorphic rocks. Sedimentary rocks result from sediment
or from transported fragments deposited in water, e.g., limestone, formed
from inorganic remains, such as shells and skeletons. Metamorphic rocks,
e.g., as marble, slate, quartzite, formed when rock masses were subjected to
high heat and pressure in geological time. Whilst the main element in cal-
cium carbonate deposits is calcium, other elements are present, magnesium
(Mg), iron (Fe), and manganese (Mn) essentially that affect whiteness, hard-
ness and specific gravity.
Typical Fillers for Polymers 55

Available in a wide range of particle sizes (range 500–1000 nm), ground


CaCO3 Ground Calcium Carbonate (GCC) is a low cost filler added to extend
and cheapen the widest range of thermoplastic polymer systems. In contrast,
synthetic calcium carbonate as obtained by precipitation, has generally a finer
and more regular particle size and therefore exhibit a semi-active character.
Ultrafine, suitably coated (for instance with calcium stearate) precipitated CaCO3
Precipitated Calcium Carbonate (PCC) can even be considered as a sophisti-
cated (and of course more expensive) filler with the following characteristics:
primary particle size (in the 20–70 nm range); narrow particle size distribution;
regular and controlled crystalline shape (rhombohedral for calcite; orthorhom-
bic for aragonite). PCC is prepared through a three steps process: first calcium
oxide is obtained by calcining crude calcium carbonate at around 900°C, then
water is added to give calcium hydroxide, and eventually carbon dioxide is
passed through this solution to precipitate the desired calcium carbonate:

CaCO3→CaO + CO2↑
CaO + H2O→Ca(OH)2
Ca(OH)2 + CO2→CaCO3↓ + H2O

The fineness of the particles, as well as the crystal form is controlled by


adjusting the temperature, the concentration of reactants and the timing of
the operations. Because it is a synthetic material, PCC offers a range of tech-
nical capabilities that are beyond the ones of GCC fillers, particularly when
ultrafine coated grades are considered. Indeed, the fabrication of synthetic
calcium carbonate through the above chain of chemical reactions is so suffi-
ciently versatile that morphology and crystalline size of PCC can be varied at
will. With respect to potential uses in polymer technology, ultrafine coated
PCC grades exhibit several important characteristics:
1. Primary particle size in the 20–70 nm range,
2. Regularity of shape (essentially rhombohedral),
3. Narrow particle size distribution,
4. High purity, and
5. Surface coating (generally with calcium stearate).
There are in fact a number of plastic applications where PCC cannot be
challenged by any GCC grade.
Calcium carbonate, is a very common mineral, with deposits nearly every-
where on the planet and more than 60 large companies are producing and
processing it; but only a few companies have worldwide significance: Minerals
Technologies Incorporated (MTI), by far the largest with around 55 operat-
ing units worldwide, Omya, the largest producer of ground calcium carbon-
ate in the world, is the second, Imerys has eight plants, Solvay (Belgium) is an
important producer and merchant of PCC in Europe, Okutama Kogy (Japan)
produces both GCC and PCC and there are also several Chinese companies
of growing importance.
56 Filled Polymers

Dolomite, named for the French mineralogist Deodat de Dolomieu, is a


natural blend of calcium and magnesium carbonate CaCO3.MgCO3 that is
commonly found in deposits of a sedimentary rock called dolostone. Two
types of materials are called dolomite, a true chemically uniform calcium
magnesium carbonate CaMg(CO3)2, and a dolomitic limestone, i.e., an irregu-
lar mixture of calcium and magnesium carbonates. Dolomite is harder and
denser than the calcite form of CaCO3 (limestone), and is more chemically
inert, namely more resistant to acid attack. When finely grinded, dolomite
can be used as an inert filler in thermoplastics.

4.2.1.5  Miscellaneous Mineral Fillers


Barium sulfate: BaSO4; specific gravity: 4.5 g/cm3; essentially used to increase
the density of polymer systems. High purity synthetic barium sulfate, often
referred as “blanc fixe” is used as filler and extender in a great variety of
plastics. Blanc fixe is precipitated with well-controlled particle diameter, in
the micron range, from pure and filtered barium salt and sodium sulfate
solutions. It can then be used in producing translucent plastic sheets or as an
additive, which improves mould release of injection-molded parts. Lithopone
(ρ = 4.2 g/cm3) is a nontoxic white pigment made by coprecipitating BaSO4
and ZnS, which can be used for the same purpose, providing the zinc sulfide
does not interfere with certain chemical aspects of the polymer material.
Zinc oxide: ZnO; not considered as a filler for polymers but used in the
rubber industry for its major role in the complex sequence of chemical reac-
tions that are involved in bridging elastomer chains (so-called vulcaniza-
tion). Typical loading is 5 phr, always associated with 2–3 phr of stearic acid
that is reacting with ZnO to yield zinc stearate, readily soluble in the com-
pound, at the vulcanization temperature.
Magnesium oxide: MgO; used as chemical buffer of reaction products that
have an acidic character. For instance, MgO is a necessary ingredient in the
curing formulation of fluoroelastomers where it captures the fluorohydric
acid (HF) resulting from chains bonding.
Calcium oxide: CaO; used as a moisture absorber in certain polymer
materials.
Table 4.8 compares the properties of a few selected white minerals used as
fillers in the rubber and thermoplastic industries.

4.2.2  Silica Fabrication Processes


4.2.2.1  Fumed Silica
So-called fumed silica results from a pyrolitic process, as the product of the
hydrolysis of silicon tetrachloride vapor in a flame of hydrogen and oxygen,
according to the following set of chemical reactions:

2 H2 + O2→2 H2O
SiCl4 + 2 H2O→SiO2↓ + 4 HCl↑
Table 4.8
Typical White Minerals Used as Fillers for Polymers
Specific Gravity Surface Area Specific. Heatb Cp Thermal Conduc­tivityb
Filler Chemical Composition (g/cm3) (m2/g) Mohs Hardness (at 298 K)(J/g.K) (W/m.K)
Clay Kaolinites Al2O3.2SiO2.2H2O 2.60 10–30a 2.5 93–102 2.0–3.2
Typical Fillers for Polymers

Talc 3MgO.SiO2.H2O 2.70–2.80 2–12a 1 80–85 3.0–5.3


Mica KAl2(AlSi3O10)(F,OH)2 2.75–2.90 19–140a 2.0–2.5 100 2.2–2.3
K(Mg,Fe)3AlSi3O10(F,OH)2
Quartz (grinded) SiO2 2.65 n.a. 7.0 69–72 6.5
Wollastonite CaSiO3 2.87–3.09 3–100a 4.5–5.0 62–73 n.a.
Silica (synthetic) SiO2 2.65–2.70 100–240 7.0 44.7 21
Calcium carbonate
(synthetic) CaCO3 2.70–2.72 3–12 3 78–82 3.5–5.0
Dolomite (grinded) CaCO3.MgCO3 2.40–2.90 5–20a 3.5–4.0 80–90 4.9–5.5
Barium sulfate
(blanc fixe) BaSO4 4.3–4.5 7–20 3 0.44 1.3–1.5
a Specific surface area data were compiled from various sources and may vary considerably, outside the quoted range, depending on the milling/grind-
ing process.
b Thermal data were obtained from various sources of geological information; the actual data of the mineral used as filler in the polymer industry might

be slightly different.
57
58 Filled Polymers

Overall reaction:

SiCl4 + 2 H2 + O2 1800



oC
→ SiO2↓ + 4 HCl↑

During the combustion process, elementary spheres of molten silica are


formed, whose diameters are the 7–30 nm range, depending of the process
parameters. Silica grades with final surface areas ranging from 400 to 100
m2/g, respectively are produced. The molten silica spheres, so-called pri-
mary particles, collide and fuse with one another to give three dimensional,
branched, chain-like objects (“string of pearls”). The largest dimensions of
these chainlike objects are in the few tenths of a micron range. As the fused,
aggregated primary spheres cool down below the fusion temperature of
silica (approximately 1710°C) further collisions occur which lead to revers-
ible mechanical entanglement or agglomeration. Further agglomeration also
takes place when silica is collected from the fumes. The residual adsorbed
HCl on fumed silica surface is then reduced to less than 200 ppm through
calcination. Certain grades go through additional densification processes
that raise the bulk density, for instance from the normal average value of 0.03
to more than 0.07 g/cm3 depending on the grade. The true density of fumed
silica is 2.2 g/cm3. As expected, the purity of the starting materials, the com-
pleteness of the pyrolitic process and the care in handling the product con-
dition the degree of chemical purity of fumed silica. Only traces of metallic
contaminants, e.g., aluminum, bore, calcium, zinc, etc., in the ppm range are
found in typical fumed silica grades. As no ionic impurities are present in
significant amount, fumed silica has excellent dielectric properties. Fumed
silica is a non-porous, amorphous (i.e., noncrystalline) material.

4.2.2.2  Precipitated Silica


Basically silica can be obtained by a wet process from solutions of alkali sili-
cate, preferably sodium silicate, from which amorphous silica is precipitated
by adding acid. After is has been filtered, washed and dried, the product con-
sists of 85–90% SiO2 and 10–15% water, either present in the material structure
(silanol groups) and/or physically bound on the surface. By controlling the
precipitation parameters, such as temperature, pH, electrolyte concentration,
stirring rate and duration, silica with different surface areas can be obtained, in
the 25–700 m2/g. Various grades of amorphous precipitated silica are obtained
from a process that essentially uses sand as starting material (Figure 4.16).
Sand and sodium carbonate are first blended then molten by fusion in a
furnace at 1400°C. The resulting vitreous silicate is dissolved under pressure
in hot water. The liquid silicate is diluted at an appropriate concentration in
water before performing the precipitation stage. Silica particles precipitate
when adding sulfuric acid (or solutions of metal salts for high performance
grades), with sodium (metal) sulfate as byproduct which is removed by fil-
tration, and washing. Silica gel formation is avoided by stirring at elevated
H2SO4

Sand Sodium
carbonate Micronizing
Precipitation

Grinding
Slurry Dry
Pelletizing
Typical Fillers for Polymers

Blender Dilution powder


tank silica
Filtration

Washing
Furnace
1400° C
SiO2 + H2O
Liquid Storage silos
silicate Liquefaction
Vitreous
silicate

H2O

Packaging and delivering


Dissolution Spraying
tank and drying
Bags Bin-bags Trucks

Figure 4.16
Manufacturing process of precipitated silica.
59
60 Filled Polymers

temperatures. Silica suspended in water is then dried by spraying and the dry
powder is further treated to give various grades with respect to application
requirements. Key steps in the process are precipitation, filtration, and spray-
ing/drying but the microscopic particle morphology is essentially determined
by the ­precipitation stage, during which controlling the composition and the
ratio of reactants, the reaction time, the temperature and the concentration
monitors silica properties. When precipitating silica, the reaction mixture
is maintained in the alkaline pH region, in order to control the aggregation
of primary particles whose diameter is larger than 5 nm. The precipitation
process affects silica characteristics such that structure (oil absorption and
compressibility), chemical properties (pH, silanol group density and optical
properties). Drying is of course a major cost component in manufacturing
precipitated silica. The most common technique is spray drying but rotary
drying is also used, which give rise to different particle shapes, degrees of
agglomeration, and to a lesser extent porosity. The dry powder silica (which
still contains 6–8% of hydrogen bonded water) can be further treated in a
number of ways (grinding, milling, micronizing) in order to reduce the size
of the clusters (agglomerates) formed during the drying process and to obtain
specific particle size distributions. Final products are stored in silos under
controlled atmospheric conditions, before packaging and delivering.
Precipitated silica is a safe amorphous (noncrystalline) mineral, an inert,
nontoxic, powder, and considered as nonhazardous with respect to manu-
facture, transportation and handling, in contrast with finely grinded natural
silica (quartz) or silicates which can give pulmonary deceases (lung tumors or
silicosis). Essentially, precipitated silica seems to consist in bulky three dimen-
sional assemblies of primary particles, approximately spherical. One may
consider that, as precipitation proceeds, elementary silica particles, whose
surface is rich in hydroxyl groups, tend to adhere each other and strongly
stick together through hydrogen bonding. Then, through packing of addi-
tional elementary particles, these first clusters grow in size higher than 4–5
nm before they further agglomerate in larger assemblies. Direct observation
of such a cluster formation mechanism is hardly possible but is easily sup-
ported by basic thermodynamic considerations. Indeed, hydrogen bonding
corresponds to an energy of around 20 kJ/mole and, during the precipiation,
the thermal energy that could conflict with the packing of elementary par-
ticles can be estimated as RT ≈ 2.69 kJ/mole , would the precipitation occurs
at 50°C. So thermal agitation cannot really hinder clustering. An additional
argument is provided when considering the time it takes for a spherical parti-
cle of diameter d to diffuse in a medium of viscosity η at temperature T, i.e.

3 π η d3  4 π  1/3 
tdiff =   − 2 (4.13)
16 kB T  3 Φ  

where kB is the Boltzmann constant (1.3 × 10 –23 J/K) and Φ the volume
fraction of particles. If one considers that the viscosity of the suspending
Typical Fillers for Polymers 61

medium (i.e., water + sulfuric acid) is around 2 × 10 –6 Pa.s, then one ­calculates
tdiff ≈ 6.3 × 10−9 s at T = 50°C. Such a diffusion time is so short that cluster-
ing of precipitating elementary SiO2 particles can be considered as a nearly
instantaneous process. Once in close contact, clustered elementary particles
will then remain so because cumulated hydrogen bonding forces between
particles is considerably larger than thermal energy. It follows that large
clusters of elementary particles of silica is expected a common observation.
Indeed, Medalia has published comparative electronic microscopy pho-
tographs of a typical assembly of carbon black particles, of fumed silica
particles and of precipitated silica particles.45 There are both similarities
and differences between carbon black and so-called high structure silica,
either fumed or precipitated. At equal aggregate/cluster cross dimension,
carbon black elementary particles are larger, and fumed silica has a more
open structure than precipitated silica. In his paper, Medalia underlines the
difficulties in obtaining such pictures: first the filler must be thoroughly
dispersed in a liquid, then the suspension is disposed on the appropriate
support and the liquid is evaporated, using special precautions to prevent
flocculation. It seems obvious that such a technique can hardly prevent the
occurrence of artefacts, if the particles stick together owing to hydrogen
bonding. It follows that if indeed aggregates are likely the ultimate form of
carbon blacks, no certainty can be established in the case of silica, except
maybe fumed silica, because the process occurs at such high temperature
that local fusion of elementary particles is plausible. The author doubts that
it is the case for precipitated silica. It is worth underlining here that, if it
is well established that the aggregate is the reinforcing object with carbon
black, there is indeed some controversy today about the actual role of silica
clusters.
The above thermodynamic considerations would lead to the conclusion
that, if only interparticle hydrogen bonds are involved in the clustering pro-
cess, then silica clusters might be considerably more brittle structure than
carbon black aggregates. Consequently the ideal (theoretical) state of dis-
persion of silica could be as well an homogenous distribution of elementary
SiO2, in sharp constrast with carbon black for which it is well established
that the reinforcing structure is the aggregate. As we will see later, with
diene rubbers, reinforcement by silica is only achieved through a chemical
modification of silica surface and the strong silica-rubber interactions that
develop only during the vulcanization stage. Recent data suggest that in
an optimal dispersion stage, silica clusters could be totally dislocated with
spherical silica particles evenly distributed.46 Would it be the case, clus-
tered silica particles, maintained together in the rubber matrix by hydrogen
bonding (≈20 kJ/mole, to be compared with C–C, 347 kJ/mole and C = C,
613 kJ/mole), are indeed relatively weak structures that surely do not play
the same role as carbon black aggregates. The true meaning of the terms
“high structure silica,” as used by silica manufacturers would need there-
fore clarification.
62 Filled Polymers

Whilst there is no standard classification scheme (as for carbon blacks),


there are three main types of synthetic precipitated silica:

• Standard grades
• Grades for cosmetic applications (e.g., toothpaste), animal feeding
and human foodstuff, which require particular specifications
• High performance grades for various industrial applications, namely
elastomer reinforcement.

4.2.3  Characterization and Structural Aspects of Synthetic Silica


In sharp contrast with carbon black, there is no standard classification for
commercially available silica but elementary particle size and the morphol-
ogy of the reinforcing object (that we call the “cluster” in the case of silica,
keeping the term “aggregate” for carbon black) are recognized as two impor-
tant parameters for reinforcement performance. However, surface character-
istics play a larger role in the case of silica than for carbon black, thanks to
the rich surface chemistry of silica particles.
Depending on the manufacturer and the manufacturing process, i.e., essen-
tially by pyrogenation or by precipitation, different characteristics are mea-
sured, not always with standard methods. Primary characteristics include
(see Table 4.9 for typical ranges):

• Specific surface area: using the well known BET method for nitro-
gen adsorption (ASTM D3037/4820).
• Average cluster size (or morphology): DBP absorption is used
(ASTM D2414); TEM can be used for a more precise (but also more
tedious) measure of cluster and elementary particle sizes.
• Porosity: ideally through mercury absorption, to be preferred to
more common oil absorption techniques.
• pH: silica are normally slightly acidic (5.5–7.0).
• Purity: most synthetic silica contain more than 98% silicon dioxide,
with minute impurities such as Fe2O3 and water soluble Na2O, plus
some other elements depending on the process and or the starting
material. SiO2 content is determined gravimetrically by fuming off with
hydrofluoric acid. The sulphate and chloride contents are determined
by potentiometric titration. Minutes impurities are assessed using ana-
lytical techniques such as atomic absorption or ion chromatography.
• Specific gravity and tamped density: the specific gravity of silica is
approximately 2 g/cm3 but it is the tamped density that is the first
industrial concern with respect to material handling, health and
safety of workers. The tamped density (also called bulk density)
is calculated from the weight of a sample with respect to a given
Typical Fillers for Polymers 63

Table 4.9
Typical Property Ranges for Synthetic Silica
Characteristic Precipitated Silica Fumed Silica
Particle size (μm) 10–12 0.007–0.3
N2 specific surface area (m2/g) 110–240 100–400
Moisture content at 105°C, 1 h (max%) 5 0.5–2.0
Ignition loss at 10000C for 1 h (max%) 10 5–6
Bulk density (g/cm3) 0.08–0.40 0.05–0.2
pH 5% Aqueous suspension 6.5–7.2 3.7–4.8
Water absorption (%) 175–250 –
Oil absorption (%) 180–270 250–320
Residue on 325 mesh, wet sieving (max%) 0.3 0.5
Soluble salt (%max) 0.5 –
R2O3 (max%) 0.3 –
SiO2 (max%) 89 99.8
Specific gravity (g/cm3) 1.95 2.0
Refractive index 1.46 –

volume, and is reported in g/dm3 or in g/cm3. Final process treat-


ments, such as grinding, milling, and micronisation allow to some-
what control the tamped density:
• Between 0.26 and 0.36 g/cm3 for so-called micropearl silica
• Between 0.1 and 0.3 g/cm3 for powder silica
• Between 0.26 and 0.36 g/cm3 for milled silica
• Between 0.05 and 0.1 g/cm3 for micronised silica

• Moisture and ignition loss: depending on the manufacturing process,


silica contain both physically and chemically bound water (obviously
more for precipitated silica). The physically bound water, generally
referred as “drying/moisture loss,” is released by heating to 105°C
for two hours (usually in the range of 5–8 wt%) whereas dehydration
of surface silanol groups occurs at high temperatures (between 200
and 1000°C). Bonded water in the form of silanol groups, generally
referred as “ignition loss,” is removed after two hours of ignition time
at 1000°C. The ignition loss is determined on samples that are previ-
ously dried for two hours at 105°C. With respect to their respective
moisture content, precipitated and fumed silicas are also referred by
some authors as hydrated and anhydrous silicas, respectively.

In addition to the above primary physical characteristics, there is a trend


(at least in the R&D literature) to characterize the surface activity of silica in
terms of specific components of the adsorption energies of chemicals with
64 Filled Polymers

known polarities. IGC techniques are used (see Section 4.1.5) but the informa-
tion is available only for a few commercial grades (see below) Manufacturers
are practically the only source of characterization data for (commercial) silica
grades but how the structure of silica affect the properties of polymer com-
pounds remains poorly documented, essentially because few test results can
be considered for analysis. Amongst the various reasons that can be offered
for this situation, there is surely the fact that, compared with carbon black,
there are no simple and reliable test methods for directly characterizing the
structure of silica. In addition there is nearly no availability for (fumed or
precipitated) silica with similar surface areas but different structures.
Table 4.10 gives typical data as provided by several suppliers of silica. In
their technical data sheets, nearly all manufacturers describe their prod-
ucts essentially with respect to the specific surface area assessed using the
BET method, the percent weight loss after heating for 1–2 hours at 105°C,
the pH of a 5% suspension in water and the product form. Most producers
also specify the SiO2 content, generally above 98% for the precipitated silica
and above 99.8 for the fumed silica; a few ones give the tapped density and
typical particle dimensions. In 1976, with respect to the already established
ASTM classification for carbon blacks; Wagner47 suggested a letter/number
classification schema for silica with a letter, either “A” or “H” to reflect the
production method, the former for fumed silicas because they are relatively
“anhydrous,” the latter for precipitated silicas, which have around 5% bound
and adsorbed water. Silicas prepared through other processes would receive
the heading letter “C.” A first digit would refer to the elementary particle
size group and two last digit would be arbitrarily assigned but with respect
to the complexity of the structure. This very clever proposal by Wagner has
unfortunately not (yet) been adopted, despite its obvious advantages.
It must be noted however that, thanks to their very rich surface chemistry,
silica are prone to relatively easy chemical modification, either using the vast
possibilities of silane chemistry or more generally all the condensation reac-
tions that organic chemistry can offer with hydroxyl groups. Quite a large
number of so-called “modified” silicas (and silicates) are consequently avail-
able, not all however for polymer applications, but whose detailed descrip-
tion is outside the scope of the present book. Chemically modifying silica
has a cost however, sometimes excessive when organic solvents have to be
eliminated. Therefore, we will be pay attention to modified silica only in the
(rare) case of real advantages in significant industrial applications. In rubber
technology, when used in nonpolar elastomers, silica must be treated with
a silane, but despite early attention paid for presilanated products, the mas-
sive use of silica (in tire technology) developed when it became practically
feasible to perform in situ silanisation of silica, i.e., during the compound-
ing operations, essentially in internal mixers. Such a development was a tre-
mendous challenge over the last quarter of the twentieth century, nowadays
globally achieved through a combination of sound engineering pragmatism
and advanced scientific research, as we will describe in Chapter 6.
Table 4.10
Silica, Suppliers’ Data
Producer Product Name Grade N2 Spec. Area (m2/g) pH Heating Loss (%) Product Form
1. Precipitated Silica
Evonik [ex Degussa] Ultrasil 880 35 10.5 5.5 Powder
(Germany) 360 50 9 5.5 Powder
AS7 60 11.5 5.5 Powder
Typical Fillers for Polymers

VN2 125 6.9 5.5 Powder


VN2 GR 125 6.9 5.5 Granules
7000 GR 175 6.8 6.5 Granules
VN3 175 6.2 5.5 Powder
VN3 GR 175 6.2 5.5 Granules
7005 185 6 5 Spherical particles
Glassven (Venezuela) Rubbersil RS-50 68 ± 22 6.0–9.0 n.a. Powder
RS-250 200 ± 10 6.0–6.8 n.a. Powder
RS-200 180 ± 10 6.0–6.8 n.a. Powder
RS-150 150 ± 10 6.5–7.3 n.a. Powder
RS-120 120 ± 10 6.5–7.3 n.a. Powder
Lanxess (Germany) Vulkasil C 50 ± 10 9.0 ± 0.7 5.5 ± 1.5 Powder
S 175 ± 20 6.2 ± 0.8 5.5 ± 1.5 Powder
S/KG 175 ± 20 6.2 ± 0.8 5.5 ± 1.5 Granules
N 125 ± 20 6.9 ± 0.7 5.5 ± 1.5 Powder
(continued)
65
66

Table 4.10  (Continued)


Producer Product Name Grade N2 spec. Area (m2/g) pH Heating Loss (%) Product Form
Rhodia (France) Zeosil 175 GR 175 6.8 6 Granules
1115MP 110 ± 15 6–7 ~8 Micropearls
115 GR 110 ± 15 6–7 ~8 Granules
1165 MP 160 ± 20 6–7 ~8 Micropearls
165 GR 170 ± 20 6–7 ~8 Granules
195 GR 180 ± 20 5.5–7 ~8 Granules
215 GR 220 ± 10 6–7 ~9 Granules
2. Fumed silica
Cabot (USA) Cab-O-sil L-90 100 ± 15 3.7–4.3 0.5 Powder
LM-130 130 ± 15 3.7–4.3 1 Powder
LM-150 155 ± 5 3.7–4.3 <0.5 Powder
LM-150D 160 3.7–4.3 <1.0 Densed powder
HP-60 200 3.8–4.3 <1.5 Powder
M-5 200 3.7–4.3 <1.0 Powder
M-5DP 200 Powder
M-5P 200 3.7–4.3 <1.0 Powder
M-7D 198 ± 2 3.7–4.3 <1.5 Densed powder
PTG 200 3.7–4.3 <1.5 Powder
MS-75D 255 3.8–4.3 <1.5 Densed powder
MS-55 255 ± 25 3.7–4.3 1.5 Powder
Filled Polymers
H-5 300 3.7–4.3 <1.5 Powder
HS-5 325 3.7–4.3 <1.5 Densed powder
EH-5 400 ± 20 3.7–4.3 <1.5 Powder
Evonik [ex Degussa] Aerosil 130 130 ± 25 3.7–4.7 <1.5 Powder
(Germany) 130V 130 ± 25 3.7–4.7 <1 Densed powder
150 150 ± 15 3.7–4.7 <1.5 Powder
150V 150 ± 15 3.7–4.7 <1.5 Densed powder
200 200 ± 25 3.7–4.7 <1.5 Powder
300 300 ± 30 3.7–4.7 <1.5 Powder
300 SP 300 ± 30 3.7–4.7 <1.5 Structure modified
Typical Fillers for Polymers

380 380 ± 30 3.7–4.7 <2 Powder


67
68 Filled Polymers

4.2.4  Surface Energy Aspects of Silica


The surface chemistry of silica is richer and more clearly defined than that of
carbon black. As illustrated in Figure 4.17, silica surface is occupied by sizable
quantities of siloxane and silanol groups, giving rise to hydrogen interaction
with either “free” or “bound” water. Free water is easily removed by drying
at 105–250°C, while bound water is only released at 900–1000°C and results
in fact from the condensation of vicinal silanol groups. An important conse-
quence of such an oxygen rich surface chemistry is that strong inter-particles
interactions occur through hydrogen bonding. In nonpolar polymers, this
give rise to a poor dispersibility and therefore special compounding proce-
dures are needed. But suitable chemical promoters, e.g., silanes, ­easily form
covalent bonding and on one hand make mixing easier and on the other hand
participate in the reinforcing capabilities of the filler. Of course, polar poly-
mers and notably certain specialty elastomers such as polydimethylsiloxanes,
naturally interact with silica, the latter thanks to their similar chemistry.
Such a chemistry reflects in the surface energy as it can be estimated
through IGC, whose results can be conveniently expressed in terms of dis-
persive γ s and polar γ s components. As we have seen, carbon blacks exhibit
d p

a high dispersive component, in the 100–400 mJ/m2 range (at 150°C), actually
proportional to their specific surface area, and a relatively low polar compo-
nent, nearly constant, whatever the grade. On the reverse, reinforcing silica
grades have a lower dispersive component (in the 20–60 mJ/m2 at 150°C),
but a higher γ sp . This explains why filler–filler interaction dominates in silica

H H
OH
O O OH OH OH OH OH O

Si Si Si Si Si Si Si Si Si

O O O O O O O
O O O O O O O O O O O O O
Free silanol Silanediol Siloxane
Vicinal silanols Silanetriol
=> hydrogen bonding
between particles (rare, if any)
Surface density : 2 to 8 OH per nm2 (depending on specific area and production process)
Strong reactivity with water (moisture) :
H
H H
O O
H H H H H H +H O H
+H2O 2 H
H
O O O O O O -H2O O O
-H2O
Si Si Si Si Si Si Si Si Si Si

O O O O O O O O O O O O O O OO O O O O OO OO O O

Figure 4.17
Surface chemistry of silica.
Typical Fillers for Polymers 69

filled compounds, thus leading to the formation of silica networks and the
associated difficulties in dispersing such a material in nonpolar polymers.
Dispersive and polar components would therefore appear as a key infor-
mation regarding the surface activity of silica, unfortunately available only
through research literature on a few commercial grades, and with differ-
ences in tests conditions as used by the authors. Table 4.11 gives published
data on commercial silica grades, as compiled in the literature.
Anhydrous, fumed, silicas have a higher γ s than hydrated, precipitated
d

silicas and, moreover, the lower the specific surface area, the lower the dis-
persive component, irrespective of the production process. According to
Wang et al.,25 the difference in the dispersive component between both types
of synthetic silica might depend on the surface topology of their particles,
and less on the concentration of hydroxyl groups.
In contrast with carbon black, the primary silica particles (that cluster in
“string of pearls” or other tri-dimensional complex structure) is the small-
est dispersible unit and, therefore, the likely reinforcing entity. Strongly
clustered elementary silica spherical particles can be separated only when
exerting dispersive forces larger than the sum of all interparticle hydro-
gen bonds. Indeed, owing to their silanol rich surface, silica particles easily
assemble through hydrogen bonding to form clusters, leading eventually to
a relatively soft network, detrimental to reinforcement. Therefore one needs
to chemically promote the rubber–filler interaction by using bi-functional
chemicals and in the meantime to shield the surface of silica particles in
order to decrease interparticle interactions. Such bi-functional promoters, on
one side, react with silanol groups that densely cover the surface of silica
particles and, on the other side, will undergo covalent bonding with rubber
chains during vulcanization. When compared with carbon black, the rein-
forcement with silica proceeds thus from a totally different concept, with a
strong role played by chemistry.

4.3  Short Synthetic Fibers


There are many fibers of synthetic origin, most of them spun or extrusion-
drawn semicrystalline polymers (e.g., polyolefins, polyamides, polyesters),
which are not used as fillers for polymers. The main reason is that, for a
material to play a potential reinforcement role as a filler for polymer, large
differences in certain key properties must exist between the filler and the
polymer matrix. It follows that only three types (or classes) of fibrous prod-
ucts can be considered as valid short synthetic fibers candidates for polymer
reinforcement: glass fibers, carbon fibers and aramid fibers.
Glass fibers of control diameter (in the 10 µm range) are produced by melt
spinning techniques, essentially by pushing a molten glass of appropriate
Table 4.11
70

Surface Energy Components for Silica


N2 (BET) Specific Area Dispersive Component Polar Component γ p IGC Test
s
Silica grade Type (m²/g) γ ds (mJ/m2) (mJ/m2) Conditions Source
®
Ultrasil VN2 p 134 21.3 73.1 * 25
Ultrasil® VN2 p 134 22.9 64.0 * 29
Ultrasil® VN3 p 181 27.5 79.6 * 25
Ultrasil® VN3 p 181 34.3 71.9 * 29
Aerosil® 130 f 128 27.3 48.1 * 25
Aerosil® 130 f 128 30.7 45.4 * 29
Aerosil® 200 f 189 40.9 57.4 * 25
Aerosil® 200 f 189 44.3 55.1 * 29
Aerosil® 300 f 300 55.0 51.0 * 28
Cab-O-Sil® L90 f 90 35.0 127.0 ** 56
Cab-O-Sil® LM130 f 122 42.0 120.0 ** 56
Cab-O-Sil® LM150D-T f 154 59.0 144.0 ** 56
Cab-O-Sil® LM150 f 167 58.5 142.0 ** 56
Cab-O-Sil® LM150D-B f 183 58.0 147.0 ** 56
Cab-O-Sil® M7D f 196 56.0 149.0 ** 56
Cab-O-Sil® M5 f 209 55.5 144.0 ** 56
Cab-O-Sil® HP60 f 216 59.0 149.0 ** 56
Cab-O-Sil® M75D f 258 80.0 150.0 ** 56
Cab-O-Sil® HS5 f 285 81.0 154.0 ** 56
Cab-O-Sil® S17D f 407 86.0 167.0 ** 56
Cab-O-Sil® EH5 f 417 94.5 157.0 ** 56
* At 150°C with heptane for dispersive component; benzene for polar component.
** At 120°C with an alkane for dispersive component; THF for polar component.
Filled Polymers

  Ultrasil and Aerosil are Evonik trade marks for precipitated and fumed silica; Cab-O-Sil is Cabot trade mark for fumed silica.
Typical Fillers for Polymers 71

composition through a platinum crown with small holes. Continuous fibers


with diameters in the 5–24 µm range are drawn, cooled down and chopped
(cut) to the desired length. Depending on the composition, various type of
glass fibers are produced: E-glass (i.e., electrical grade glass), S-glass (high
stiffness), C-glass (chemical grade, for its best resistance to chemical attack),
R-glass (high strength, special grades, mainly for aerospace glass-reinforced
applications). E-glass is an alumino-borosilicate, low alkali composition
(SiO2, 54%; Al2O3, 14%; CaO + MgO, 20–22%; B2O3, 10%; Na2O + K2O, < 2%),
which gives excellent fiber forming capabilities. It is therefore a high
­production rate, low cost grade, almost exclusively used as reinforcing mate-
rial. S-glass (SiO2, 65 wt%;.Al2O3, 25 wt%; MgO,10 wt%) is a higher stiffness
grade, more difficult to produce than E-glass and therefore more expensive.
It is used only when high mechanical properties are required. C-glass (SiO2,
64.6%; Al2O3 Fe2O3, 4.1% CaO, 13.4%; MgO, 3.3%; Na2O·K2O, 9.6%; B2O3 BaO,
5.0%) is also a fiber reinforcement grade, made and applied specifically for
high chemical resistance. The most frequently used short glass fibers (SGF)
for filling (thermoplastic) polymers are E-glass and S-glass grades. Such
grades have typical diameters of 10 µm, with length in the 1–10 mm range;
fiber aspect ratios are thus in the 100–1000 range.
Carbon fibers combine very high tensile properties (strength and modulus)
with low weight. As the high tensile strength is maintained up to extremely
high temperatures, carbon fibers are ideal materials for lightweight struc-
tures in highly demanding applications, e.g., in aerospace. It can be con-
sidered that Thomas Edison, in the late nineteenth century, was the first to
produce carbon fibers as filament for incandescent lamps. But the first com-
mercial carbon fibers were not really produced until the early 1960s.48 The
basic principle is first to manufacture a carbon rich filament, then to carbon-
ize it at very high temperature (above 2000°C) under controlled atmosphere
conditions, in order to eliminate noncarbon atoms. A number of different
processing techniques as well as different precursor routes were developed,
but the most important among them are polyacrylonitrile (PAN-fibers) and
mesophase-pitch-precursor (MPP-fibers). Carbon fiber filaments have diam-
eters in the 5–10 µm range but are produced as yarns of several thousand fil-
aments. Short carbon fibers can be produced by chopping with fibers aspect
ratio in the 100–1000 range.
Aramid fibers, best known under the commercial trade name Kevlar®, are
polypara-phenylene terephtalamide fibers first commercialized by Dupont
in 1971, that combine high strength and flexibility with light weight. A
similar product with roughly the same chemical structure was introduced
by Akzo in 1978 under brand name Twaron® and is now manufactured by
Teijin. Spun as long fibers, typically yarns of 700–1000 filaments of around
12 µm diameter, Kevlar was first used as a replacement for steel wires in rac-
ing tires, then was used to make tissues and fabric sheets that are nowadays
used as such or as a component in multilayer composites. Short aramid fibers
are produced either as pulp (a highly fibrillated form), or as floc (precision
72 Filled Polymers

cut short fibers less than 1 mm in length), or as staple (precision cut short
fibers of 7–10 mm length), to be used as a fibrous filler for polymers. Short
aramid filaments are thus available with fiber aspect ratios in the 80–800
range.
Table 4.12 gives a few typical properties of synthetic fibers, to be com-
pared with similar data on either thermoplastics or vulcanized elastomers
(for instance tensile strength in the 0.06–0.09 GPa and 0.02–0.05 GPa ranges,
respectively). Polyamides, polyesters, and HS polyethylene fibers are quoted
for comparison only, since their relatively low melting points limit their use
in most polymers. Aramid fibers are interesting in this respect since such
materials do not melt but start to decompose above 420°C.
Short synthetic fibers are used as reinforcing materials in general ­purposes
(e.g., PP, PE) and technical thermoplastics (e.g., PBT, PA66, PA6/6T). Current
loadings are up to 30–40 wt%. Basically the reinforcing effect results from
the large difference in tensile properties between fibers and polymers, and
with respect to the average tensile strength and modulus of most thermo-
plastics and (vulcanized) elastomers, short glass, aramid, and carbon fibers
are the most interesting synthetic materials for polymer reinforcement
(Figure 4.18). Good adhesion between the polymer matrix and the fibers
is generally required, and is usually obtained through a chemical process:
fibers are treated with chemical agents that promote fiber-to-polymer adhe-
sion. However improvements in mechanical properties are also markedly
depending on fibers’ orientation, with respect to the principal strain direc-
tion, as we shall see in detail later.

4.4  Short Fibers of Natural Origin


Natural fibers are structures of biological origin, which essentially consist
in cellulose, hemicellulose and lignin, and smaller quantities of extractible,
­proteins and other inorganic elements. The exact composition depends of
course of the very origin (i.e., plant, crop, tree) and the treatment. Depending
of the origin, one may distinguish seed fibers (cotton, kapok), leaf fibers
(sisal, agave), fruit fibers (coconut), stalk fibers (bamboo, grass, wood flour).
From a biological point of view, there are many classes of natural fibers but
not all are of interest as fillers for polymers. So-called hard fibers account
for approximately 90% of the world production, but only a small propor-
tion is used to manufacture composites with polymers. With respect to the
scope of the book, we pay attention only to short natural fibers, i.e., either
chopped long fibers or flour obtained through the mechanical reduction
of bulk natural products, e.g., wood to their fibrous component. A num-
ber of short natural fibers used to prepare polymer composites are byprod-
ucts of various industries (e.g., textile, wood, etc.). Properly dispersed in a
Table 4.12
Typical Properties of (Short) Synthetic Fibers; Compiled from Various Sources (Manufacturers’ Data Sheets, Scientific Publications)
Specific Gravity Tensile Strength Tensile Modulus Elongation at Poisson’s Maximum Working
Fiber Type (g/cm3) (GPa) (GPa) Break (%) Ratio (v) Temperature (°C)
Typical Fillers for Polymers

E-glass 2.54–2.60 2.0–3.8 73–81 4.0–4.9 0.17–0.27 350–400


S-glass 2.48–2.49 4.4–4.8 88–91 5.4–5.8 0.22–0.27 350–400
C-glass 2.62–2.63 3.1–3.8 80–81 4.6–4.6 0.24–0.27 350–400
Carbon 1.75–2.10 1.9–5.7 240–531 0.3–2.1 0.23–0.28 500–600
Aramids 1.44–1.45 2.9–3.8 60–136 2.4–3.6 0.36 300–310
Steel wire 7.65–7.80 2.0–2.8 197–200 2 0.28–0.30 300–400
Polyamides 1.10–1.16 0.7–1.0 6–7 18.3–26.0 0.33 120
Polyesters 1.37–1.38 0.6–1.2 14 14.5–23.0 0.35 150
High strength PE 0.96–0.97 2.6 117 3.5 0.38 60
73
74 Filled Polymers

30

20

20 Ultra-drawn
PE
20

20 Aramids
Specific tensile strength, 106 cm

20
S-glass
18

16
Carbon
14 C-glass

12 E-glass
10

8 Polyamides
Polyesters
6

4 Graphite
Steel (high modulus)
2 wires

0
0 2 4 6 8 10 12 14 16 18 20 22
Specific modulus, 108 cm

Figure 4.18
Considering synthetic fibers in terms of specific tensile strength and modulus; specific proper-
ties are obtained by dividing the tensile property by the product (density times the gravitation
constant g = 9.807 m/s2).

(molten) polymer, such fibers can be considered as fillers for polymers and
the resulting mixtures as polymer composites such that, during process-
ing operations, both the fibers and the polymer matrix will flow together.
It follows that polymer based products such medium-density fiberboard
(MDF, or Craftwood), whilst made with a resin binder, are excluded from
our subject.
The cellulose/hemicelluloses/lignin ratio exhibits large variation in natu-
ral fibers, depending on the vegetal source, but all fibers can be viewed as
multilayer composites of a rigid cellulose structure embedded in a lignin
matrix, with covalent bonding between both.49 Cellulose is largely crystal-
line and organized in microfibrils with typical diameters in the 10–30 nm
range, with a supramolecular helicoidal structure that gives it an elasticity
modulus of around 136 GPa, i.e., twice the value of glass fiber. Hydrogen
bonding along and between cellulosic chains plays a role in the mechanical
Typical Fillers for Polymers 75

properties but gives rise also the large variability of physico-chemical


properties of natural fibers. Misnamed hemicellulose is in fact a group
of polysaccharides (five to six carbon ring sugars) with a large degree of
branching and therefore an amorphous nature. Very hydrophilic, it forms
the supportive matrix in which cellulose microfibrils are embedded. Lignin
is a complex hydrocarbon with both aliphatic and aromatic components of
very high molecular weight. Totally amorphous and hydrophobic, lignin
provides rigidity to plants and hence natural fibers. Up to five hydroxyl
and methoxyl groups, plus a few carbonyls have been identified in lignin,
making it soluble in hot alkali, easily oxidized and very reactive with phe-
nols The initial cellulose/lignin ratio can be changed through suitable
(chemical) treatments and both cellulose and lignin are organic polymers,
the latter being more hydrophobic (because of its aromatic nature) than the
former. Generally, tensile strength and modulus of natural fibers increase
with increasing cellulose content. Thermal degradation is a limiting factor
in using natural fibers as fillers for polymers, as most fibers start loosing
stiffness at around 160°C; lignin starts degrading at around 200°C (with
oxidative conditions accelerating the process) but destruction of cellulose
crystalline order occurs only above 300°C.50 Thermal conductivity of natu-
ral fibers is weak, of the order of 0.12–0.22 W/m.K, again a limiting factor for
polymer-like processes. And all fibers, not only exhibit dimensional varia-
tions when exposed to water but tend also to retain moisture, with asso-
ciated variations in thermal and mechanical properties. All those factors
severely limit the use of natural fibers as fillers for polymers and, as a rule
of thumb, all polymers whose melting range is above 180°C are ill suited
for preparing natural fibers filled composites. Table 4.13 gives some typical
characteristics of a few selected natural fibers of interest for applications in
polymer composites49,51–54
Most natural fibers need special preparation processes before they can
be used in polymer composites. First, after a sufficient drying of the raw
material, there is a mechanical or thermo-mechanical defibrillation step (or
grinding in the case of wood flour), followed by chopping to the (average)
desired length. Then sorting and screening are made, in order to remove use-
less solid contaminants, for instance minerals and dried proteinic residues,
and to achieve a somewhat controlled fibers size distribution. Depending
on the fiber, such operations can be more or less complicated and, of course,
involve some degradation, namely in what the lignin part is concerned.
Natural fibers often suffer from a large degree of nonuniformity in their
characteristics: chemical composition, crystallinity, surface properties,
diameter, cross-sectional shape, length, strength, and stiffness.51 Another
important issue in using natural fibers in polymer composite is moisture
absorption, which generally degrades mechanical properties like tensile
strength. Except wood flour in certain parts of the world, were sufficient
market developments have been achieved (for instance in USA and Canada
for so-called WPCs), constant and guaranteed material specifications for
76

Table 4.13
Natural Fibers as Potential Fillers for Polymers
Length [Typical] Hemicellulose Tensile Modulus Elongation at
Fiber (mm) Diameter (µm) Cellulose (%wt) (%wt) (GPa) Break (%)
Jute 1.5–5 5–25 60–70 14–20 20–55 1.5–2.0
Sisal 100–125 100–400 65 12 9–22 2–14
Cotton 10–60 12–25 85–90 5.7 6–12 6–8
Flax 4–60 12–30 71 19–20 28–70 1.3–3.3
Hemp 5–40 20–50 68 15 30–70 1.7–2.7
Abaca (Manila hemp) >60 25–200 52–63 20–25 12–35 2.1–2.4
Kenaf (close to jute) 2.5–4.5 14–33 72 20 53–60 1.7–2.1
Pine (flour) 2.5–3.0 – – – 11.3–11.5 n.a.
Birch (flour) 1.0–1.3 – – – 14.5–14.7 n.a.
Beech (flour) 1.0–1.3 – – – 13.6–13.8 n.a.
Silk (Bombyx Mori) 3–9.105  * 10 – 0.65–0.75 18–20

* Typical length of the thread of raw silk from one cocoon.


Filled Polymers
Typical Fillers for Polymers 77

natural fibers are still a serious obstacle to sizeable industrial applications,


despite encouraging progresses in line with concerns about sustainable
development.
There are many parameters which may affect the performance of a given
fiber in a given polymer, and with respect to extremely large variety of
possible fiber-reinforced polymer systems, an exhaustive coverage is nearly
impossible, despite the already considerable research literature on the sub-
ject. Only a few basic trends can be identified. In addition to their own
properties and whatever is the polymer (system) considered, all natural
fibers must be properly dispersed in the matrix and adequate fiber-polymer
interfacial effects must be achieved. Indeed undispersed fiber bundles are
obviously failure initiation sites and a basic principle for fiber reinforce-
ment is that effective load transfers occur at the fiber-matrix interface, as
clearly demonstrated through micromechanical considerations (as will be
detailed in Chapter 7). It is quite logical to expect that both fiber disper-
sion and fiber-matrix interfacial effects can be improved through suitable
chemical means. Simple considerations about the basic chemical aspects
of natural fibers help understanding where and how improvements can
be expected when either chemically modifying natural fibers or using so-
called bonding systems.
As mentioned above, from a chemical point of view, cellulose, hemicellu-
lose, and lignin are, in varied proportions, the main components of ­natural
fibers, with sometimes minor quantities of heteropolysaccharides (e.g., pec-
tins), waxes (natural polyalcohols), and other chemicals. The three main
components have hydroxyl groups with varied degree of accessibility or
chemical reactivity that conditions their basic chemical properties and hence
the possibility for chemical modification or interaction. Cellulose has strong
microcrystalline regions, enhanced by internal hydrogen bonding; is resis-
tant to strong alkali, relatively resistant to oxidizing agents but easily hydro-
lyzed by acids. Hemicellulose is noncrystalline by nature, is very hydrophilic,
soluble in alkali and easily hydrolyzed by acids. Lignin is totally amorphous
and hydrophobic, cannot be hydrolyzed by acids, but is soluble in hot alkali,
easily oxidized and very reactive with a number of organic chemicals, for
instance phenols. It is quite clear that the chemical virtues of the main com-
ponents of natural fibers dictate the possible chemical modifications and/or
the choice of bonding systems.
As recently reviewed by Jacob John and Anandjiwala,49 there has been a
considerable research during the past decades with the objective to under-
stand and somewhat optimize the fiber–polymer interaction through chem-
ical means. Only a few common techniques will be screened hereafter.
There are essentially two approaches that can be used to chemically pro-
mote or enhance interactions between natural fibers and polymer matrix,
either through a modification of the fiber surface by a suitable pretreatment
or through the use of a suitable bonding system during the compounding
operations. A priori, the latter approach would be more economical than
78 Filled Polymers

Table 4.14
Chemical Approaches to Promote Natural Fiber–Polymer Matrix Interactions
Fiber Surface Modifications (Pretreatment) Bonding Systems (When Compounding)
Organic solvent extraction Maleic anhydride grafted polymers
Steam treatment    MA-g-Polypropylene
Alkali treatment (mercerization) under various    MA-g-Polyethylene
conditions (temperature, pressure, steam,…)    MA-g-Polystyrene
   MA-g-Poly(ε-caprolactone)
Esterification with various anhydrides (acetic, Resins systems :
butyric, methacrylic, etc) Phenol-formaldehyde
Resorcinol-formaldehyde
Resorcinol + hexamethylenetetramine
SiO2 + resorcinol + hexamethylenetetramine
Acetylation, propionylation, benzoylation, Commercial adhesive systems based on
stearation, cyanoethylation,… acrylics, epoxydes, urethanes or
cyanocrylates technologies
Grafting with isocyanate derivatives Bi-functional silanes:
   with acrylates or polyacrylic acids bis(triethoxysilylpropyl)tetrasulfane
   with acrylonitrile (TESPT)
g-methacryloxy-propyl trimethoxy
silane, etc.
Silanation (various organosilanes used); Thiols
most reactions occurring above 70°C e.g., 3-(trimethoxysiliyl)-1-propanethiol
Peroxide treatment (benzoyl peroxyde) Titanate based systems
Latex coating Isocyanates, diisocyanates
(Helium) plasma, γ-irradiation
Enzymatic treatment, fungal modifications

the former, and achievable during compounding operations. Table 4.14


describes several possibilities that have been frequently studied with cer-
tain types of fibers, while only a few ones have apparently lead to sizeable
industrial usages.
Various results have been reported using these approaches, certain giv-
ing significant improvement in mechanical properties, others giving very
little benefits, if any. It is fairly clear from Table 4.14 that, when chemical
reactions are expected in using these various approaches, the main target
are the hydroxyl groups of (amorphous) cellulose, hemicellulose or lignin in
order to render the fiber surface more hydrophobic. SEM microphotographs
of fractured specimen surfaces are often used to demonstrate fiber–polymer
bonding but, in the author’s opinion, such “evidence” could also at best suit
enhanced wetting of the fibers by the matrix, without necessarily the occur-
rence of chemical (i.e., covalent) bonding. The actual role of the fiber–­matrix
interface in improving certain mechanical properties remains however
hardly understood.
Typical Fillers for Polymers 79

References
1. G. Kühner. Reaction parameters and carbon black properties. Second International
Conference on Carbon Black, Mulhouse, France, Sept. 27–30, 1993.
2. BET, N2 adsorption method: S. Brunauer, P.H. Emmet, E. Teller. Adsorption of
gases in multimolecular layers. J. Am. Chem. Soc., 60, 309–319, 1938.
3. B.B. Mandelbrot. Les Objets Fractals—Forme, Hasard et Dimensions, 4th Ed.
Champs, Flammarion, Paris, France, 1995. ISBN 2-08-081301-3; translation in
English: Fractals, Form, Chance and Dimensions. W.H. Freeman & Co, Springer, The
Netherlands. ISBN 0716704730; The Fractal Geometry of Nature. W.H. Freeman &
Co, San Francisco, CA, 1986. ISBN 0716711869.
4. B.H. Kaye. Specification of the ruggedness and/or texture of a fine particle pro-
file by its fractal dimension. Powder Technol., 21 (1), 1–16, 1978.
5. A.G. Flook. The use of dilation logic on the quantinet to achieve fractal dimen-
sion characterization of textured and structured profile. Powder Technol., 21 (2),
295–298, 1978.
6. X. Bourrat, A. Oberlin, H. Van Damme, C. Gateau, R. Bachelar. Mass fractal
analysis of conducting carbon black morphology. Carbon, 26 (1), 100–103, 1988.
7. F. Ehrburger-Dolle, M. Tence. Determination of the fractal dimension of carbon
black aggregates. Carbon, 28 (2–3), 448–452, 1990.
8. C.R. Herd, G.C. McDonald, W.M.Hess. Morphology of carbon-black aggregates:
fractal versus Euclidean geometry. Rubb. Chem. Technol., 65, 107–129, 1992.
9. D. Göritz, H. Raab, J. Fröhlich, P.G. Maier. Surface surtucture of carbon black
and reinforcement. Rubb. Chem. Technol., 72, 929–945, 1999.
10. C.R. Herd, G.C. McDonald, R.E. Smith, W.M. Hess. The use of skeletonization
for the shape classification of carbon-clack aggregates. Rubb. Chem. Technol., 66,
491–509, 1993.
11. A.I. Medalia. Morphology of aggregates. VI. Effective volume of aggregates of
carbon black from electron microscopy; application to vehicle absorption and to
die swell of filler rubber. J. Colloid Interface Sci., 32, 115–131, 1970.
12. A.I. Medalia. Elastic modulus of vulcanizates as related to carbon black struc-
ture. Rubb. Chem. Technol., 47, 877–896, 1973; Effect of carbon black on dynamic
properties of rubber vulcanizates. ibid., 51, 437–523, 1978.
13. A.I. Medalia. Morphology of aggregates. I. Calculation of shape and bulkiness
factors; application to computer-simulated random flocs. J. Colloid Interface Sci.,
24, 393–404, 1967.
14. A.I. Medalia, F.A. Heckman. Morphology of aggregates. II. Size and shape factors
of carbon black aggregates from electron microscopy. Carbon, 7, 567–582, 1969.
15. P. Meakin. Formation of fractal clusters and network by irreversible diffusion-
limited aggregation. Phys. Rev. Lett., 51 (13), 1119–1122, 1983.
16. M. Kolb, J. Botet, R. Jullien. Scaling of kinetically growing clusters. Phys. Rev.
Lett., 51 (13), 1123–1126, 1983.
17. K.M. Jäger, D.H. McQueen. Fractal structures in carbon black polymer compos-
ites. Third International Conference on Carbon Black, Mulhouse, France, Oct. 2000,
117–120.
18. M. Klüppel, G. Heinrich. Fractal structures in carbon black reinforced rubbers.
Rubb. Chem. Technol., 68, 623–651, 1995.
80 Filled Polymers

19. A.I. Medalia. Effective degree of immobilization of rubber occluded within car-
bon black aggregates. Rubb. Chem. Technol., 45, 1171–1194, 1972.
20. P. Bertrand, L;T; Weng. Carbon black surface characterization by ToF-SIMS and
XPS. Rubb. Chem. Technol., 72, 384–397, 1999.
21. A.M. Gessler, W.M. Hess, A.I. Medalia. Reinforcement of elastomers with car-
bon black. Part IV. Interaction between carbon black and polymer. Plast. Rubb.
Proc., 3, 141–156, 1978.
22. E.M. Dannenberg. Primary structure and surface properties of carbon black—
Part 1: a study of four ISAF type carbon blacks of varying structure. Rubber Age,
98, 82–90, 1966.
23. E.M. Dannenberg. Primary structure and surface properties of carbon black—
Part 2 : properties of commercial carbon blacks of varying structure and particle
size. Rubber Age, 99, 81–85, 1966.
24. A.K. Manna, P.P. De, D.K. Tripathy, S.K. De, M.K. Chatterjee. Effect of ­surface
oxidation of carbon black on its bonding with epoxidized natural rubber
in the presence of silane coupling agent. Rubb. Chem. Technol., 72, 398–409,
1999.
25. M.J. Wang, S. Wolff, J.B. Donnet. Filler-elastomer interactions. Part I: silica sur-
face energies and interactions with model compounds. Rubb. Chem. Technol., 64,
559–576, 1991.
26. C. Saint-Flour, E. Papirer. Gas-solid chromatography : a quick method of esti-
mating surface free energy variation induced by the treatment of short glass
fibers. J. Colloid Interface Sci., 91, 69–75, 1983.
27. J.R. Conder, C.L. Young. Physico-chemical Measurements by Chromatography.
J. Wiley & Sons, New York, NY, 1979.
28. E. Papirer, H. Balard, A. Vidal. Inverse gas chromatography: a valuable method
for the surface characterization of fillers for polymers (glass fibres and silicas).
Eur. Polym. J., 24 (8), 783–790, 1988.
29. M.J. Wang, S. Wolff, J.B. Donnet. Filler-elastomer interactions. Part III: carbon
black surface energies and interactions with elastomer analogs. Rubb. Chem.
Technol., 64, 714–736, 1991.
30. H. Darmstadt, C. Roy, S. Kalliaguine, H. Cormier. Surface energy of commer-
cial and pyrolytic carbon blacks by inverse gas chromatography. Rubb. Chem
Technol., 70, 759–768, 1997.
31. H. Darmstadt, N.Z. Cao, D.M. Pantea, C. Roy, L. Sümmchen, U. Roland, J.B.
Donnet, T.K. Wang, C.H. Peng, P.J. Donnelly. Surface activity and chemistry of
thermal carbon blacks. Rubb. Chem. Technol., 73, 293–309, 2000.
32. A. Schröder, M. Klüppel, R.H. Schuster. Oberflächenaktivität von Furnacerußen.
I. Bestimmung der Oberflächenrauheit mittels statischer Gasadsorption,
Monolagenbereich. Kautsch. Gummi. Kunstst., 52, 814–822, 1999.
33. A. Schröder. Charakterisierung verschiedener Rußtypen durch systemati-
sche statische Gasadsorption; Energetische Heterogenität und Fraktalität der
Partikeloberfläche. PhD Thesis, University of Hannover, Germany, 2000.
34. A. Schröder, M. Klüppel, R.H. Schuster. Characterisierung der
Oberflächenaktivität. II. Bestimmung der Oberflächenrauheit von Furnacerußen
mittels statischer Gasadsorption, Multischichtenbereich. Kautsch. Gummi.
Kunstst., 53, 257–265, 2000.
35. A. Schröder, M. Klüppel, R.H. Schuster, J. Heidberg. Energetic surface hetero-
geneity of carbon black. Kautsch. Gummi. Kunstst., 54, 260–266, 2001.
Typical Fillers for Polymers 81

36. T. Zerda, W. Xu, H. Yang, M. Gerspacher. The effects of heating and cooling rate
on the structure of carbon black particles. Rubb. Chem. Technol., 71, 26–37, 1998.
37. M. Gerspacher, C.P.O’Farrell. Tire compound materials interactions. Kautsch.
Gummi. Kunstst., 54, 153–158, 2001.
38. A.L. Barabasi, M. Araujo, H.E. Stanley. Three-dimensional Toom model: con-
nection to the anisotropic Kardar-Parisi-Zhang equation. Phys. Rev. Lett., 68,
3729–3732, 1992.
39. G. Heinrich, M. Klüppel. A hypothetical mechanism of carbon black formation
based on molecular ballistic deposition. Kautsch. Gummi. Kunstst., 44, 419–423,
1991.
40. M. Pontier-Johnson. Noir de carbone au four : mécanismes de formation des
particules. PhD Thesis, University of Haute Alsace, Mulhouse, France, 1998.
41. M. Pontier-Johnson, J.B. Donnet, T.K. Wang, C.C. Wang, R.W. Locke, B.E. Brinson,
T. Marriott. A dynamic continuum of nanostructured carbon in the combustion
furnace. Carbon, 40 (2), 189–194, 2002.
42. A. Vidal, S.Z. Hao, J.B. Donnet. Modification of carbon black surfaces—effects
on elastomer reinforcement. Kautsch. Gummi. Kunstst., 54, 159–165, 2001.
43. P. Liu. Polymer modified clay minerals: A review. Appl. Clay Sci., 38, 64–76, 2007.
44. M.R. Mushack, A.W.Backmann. Neuburg silica: a natural functional filler. Intern.
Polym. Sci. Technol., 23 (9), 5–10, 1996.
45. A.I. Medalia. Filler aggregates and their effects on dynamic properties of rub-
ber vulcanizates. (See Figure 2, p. 63). In International CNRS Colloquium, Le
Bischenberg-Obernai, France, Sept. 24–26, 1973. CNRS, Paris, France, 1975.
ISBN 2-222- 01749-1.
46. D. Göritz. Carbon Black und Silica : Gemeinsamkeiten und Unterschiede. 8th Fall
Rubber Colloquium, DIK, Hannover, Germany, Nov. 26–28, 2008. Communication
nr 55.
47. M.P. Wagner. Reinforcing silicas and silicates. Rubb. Chem. Technol., 49, 703–774,
1976.
48. R.J. Young, R.J. Day, M. Zakikhani. The structure and deformation behaviour
of Poly(p-phenylene benzobisoxazole) fibres. J. Mater. Sci., 25 (1A), 127–136,
1990.
49. M. Jacob John, R.D. Anandjiwala. Recent dvelopments in chemical modification
and characterization of natural fiber-reinforced composites. Polym. Compos., 29
(2), 187–207, 2008.
50. D-Y Kim, Y. Nishiyama, M. Wada, S. Kuga , T. Okano. Thermal decomposition
of cellulose crystallites in wood. Holzforschung, 55 (5), 521–524, 2001.
51. A.K Bledzki, J.Gassan. Composites reinforced with cellulose based fibres. Prog.
Polym. Sci., 24, 221–274, 1999.
52. R.M. Rowell, R.A Young, J.K. Rowell. Paper and Composites from Agro-based
Resources, R.M. Rowell, Ed. Lewis Publishers, Boca Raton, FL, 1997.
53. P. Jodin. wood: Engineering material. Arbolor, Nancy, France, 1994. ISBN-10:
2907086073.
54. E. Spaˉrniņ š. Mechanical properties of flax fibers and their composites. Licentiate
Thesis, Luleå Univ. Technol., Sweden, 2006.
55. M.J. Wang. Effect of filler-elastomer interaction on tire tread performance. Part
III. Kautsch. Gummi, Kunstst., 61 (4), 159–165, 2008.
56. M.J. Wang, M.D. Morris, Y. Kutsovsky. Effect of fumed silica surface area on
silicone rubber reinforcement Kautsch. Gummi, Kunstst., 61 (3), 107–117, 2008.
82 Filled Polymers

Appendix 4

A4.1  Carbon Black Data


A4.1.1  Source of Data for Table 4.5
1. http://www.degussa.com: Data on Corax carbon blacks.
2. http://www.columbianchemicals.com: Data on rubber blacks.
3. A.I. Medalia, F.A. Heckman. Morphology of aggregates. II. Size and
shape factors of carbon black aggregates from electron microscopy.
Carbon, 7, 567–582, 1969.
4. A.I. Medalia. Morphology of aggregates. VI. Effective volume of
aggregates of carbon black from electron microscopy; application to
vehicle absorption and to die swell of filled rubber. J. Colloid Interf.
Sci., 32, 115–131, 1970.
5. P.A. Marsh, A. Voet, T.J. Mullens, L.D. Price. Quantitiative microgra-
phy of carbon black microstructure. Carbon, 9, 797–805, 1971.
6. K.A. Burgess, C.E. Scott, W.M. Hess. Vulcanizate performance
as a function of carbon black morphology. Rubb. Chem. Technol., 44,
230–248, 1971.
7. J.B. Donnet, A. Voet. Carbon Black. Marcel Dekker, New York, NY,
1976.
8. G.R. Cotten. Influence of carbon black on processability of rubber
stocks II. Extrusion shrinkage. Intern. Rubb. Conf. RUBBERCON ‘77,
paper 31, 1–13, 1977.
9. G.R. Cotten, J.L. Thiele. Influence of carbon black on processability
of rubber stocks III. Extensional viscosity. Rubb. Chem. Technol., 51,
749–763, 1978.
10. A.M. Gessler, W.M. Hess, A.I. Medalia. Reinforcement of elasto-
mers with carbon black. I. The nature of carbon black. Plast. Rubb.:
Processing, 3 (1), 1–13, 1978.

Note: All appendices were made with the calculation software MathCad® 8.0 (MathSoft
Inc., now part of PTC, Parametric Technology Corporation). Reproduced as such in the
software, active worksheets would be obtained and would give the same results providing
the layout is strictly respected since the software operates the mathematical formulas from
left to right, from top to bottom.
Typical Fillers for Polymers 83

11. A.I. Medalia. In Carbon Black—Polymer Composites. E.K. Sichel Ed.


Marcel Dekker, New York, NY, 1981.
12. G.R. Cotten. Mixing of carbon black with rubber. III. Analysis of
the mixing torque curve. Kautsch., Gummi, Kunstst., 38 (8), 705–709,
1985.
13. ASTMD-1765-86a. Carbon blacks used in rubber products.
14. A.C. Patel, K.W. Lee. Characterizing carbon black aggregate via
dynamic and performance properties. Elastomerics, 122 (3), 14–18,
1990.
15. A.C. Patel, D.C. Jackson. Carbon black characterization—Part 1:
Effects of dynamic parameters on the behaviour of carbon black in
rubber. J.M. Huber Corp. (data presented on a poster at IKT’91, June
25, 1991, Essen, Germany).
16. M-J. Wang, S. Wolff, J.B. Donnet. Filler-elastomer interactions. Part
III. Carbon black surface energies and interactions with elastomer
analog. Rubb. Chem. Technol., 64, 714–736, 1991.
17. F. Ehrburger-Dolle, S. Misono. Characterization of the morphology
of rubber grade carbon blacks by thermoporometry. Carbon, 30 (1),
31–40, 1992.
18. C.R. Herd, G.C. McDonald, W.M. Hess. Morphology of carbon black
aggregates: Fractal versus Euclidean geometry. Rubb. Chem. Technol.,
65, 107–129, 1992.
19. E. Custodero. Caractérisation de la surface de noirs de carbone; nou-
veau modèle de surface et implication pour le renforcement. PhD
Thesis, University of Haute-Alsace, Mulhouse, France, 1992.
20. M-J. Wang, S. Wolff, E-H. Tan. Filler-Elastomer interactions. Part
VII. Study on bound rubber. Rubb. Chem. Technol., 66, 163–177,
1993.
21. M-J. Wang, S. Wolff, E-H. Tan. Filler-Elastomer interactions. Part VIII.
The rôle of the distance between filler aggregates in the dyanmic
properties of filled vulcanizates. Rubb. Chem. Technol., 66, 178–195,
1993.
22. T.C. Gruber, C.R. Herd. Anisometry measurements in carbon
black aggregate populations. Rubb. Chem. Technol., 70, 727–746,
1997.
23. A. Schröder. Charakterisierung verschiedener Rußtypen durch
systematische statische Gasadsorption; Energetische Heterogenität
undFraktalität der Partikeloberfläche. PhD Thesis, Universität
Hannover, Germany, 2000.
24. A. Weigert. Rastertunnelspektroskopie an Füllstoffrußen. PhD
Thesis, Universität Regensburg, Germany, 2005.
84 Filled Polymers

A4.1.2 Relationships between Carbon Black Characterization Data

1.4

1.2
Compressed DBPA, dm3/kg

1.0

0.8

0.6

0.4

0.2

0.0
0.0 0.5 1.0 1.5 2.0
DBP absorption, dm3/kg

300

250
N2 adsorption, m2/g

200

150

100

50

0
0 50 100 150 200 250 300
I2 adsorption, mg/g

1.4
Comp. DBP absorption, dm3/kg

1.2

1.0

0.8

0.6

0.4

0.2

0.0
0 20 40 60 80 100 120 140 160
CTAB adsorption, m2/g
Typical Fillers for Polymers 85

A4.2  Medalia’s Floc Simulation for Carbon Black Aggregate


Medalia’s floc simulation data  [Source: A.I. Medalia, J. Colloid Interf. Sci.,
(Table II) 24, 393–404, 1967.]
Number Area of Particle projected area is expressed in “square
of particles a particle units”; the area Ap of a particle is 78.5 square
units
5 381 
 10 572 
Ap: = 78.54

 15 803  π.d 2 4 . Ap
  One has : Ap = = > d: = d = 10
 25 1265  4 π
 50 2145 
  Medalia has thus considered flocs of particles
 74 2970  with diameter equal to 10 units
 25 1386 
  π ⋅ d2
 74 28933 Ap := Ap = 78.54
 50 2453  4
 
 50 2497  A: = sort (Data<1>) Np: = sort (Data<0>)
 15 605 
  Medalia published Correct exponent
 15 1089  (from nonlinear fitting of
5 337  scaling exponent Medalia’s data)
Data: =  
 10 572  εM: = 0.87 εC: = 0.847
 25 1265 
  AM : = Ap ⋅ N p M
ε
AC : = Ap ⋅ N p εC
5 1
331
 10 532 
 4000
 25 1329 
 
Floc projected area, square unit

5 357 
 10 635  3000
 
 25 1236 
 50 2301  2000
 
5 291 
 10 496  1000
 
 25 1289 
 50 2214 
0
0 20 40 60 80
Number of particle (of d = 10 unit)
Using the published value
Medalia's data
of 0.87 for the scaling expo- Published exponent
nent leads to an overestima- Correct exponent
tion of the projected area;
with ε = 0.847 the ­estimation
is correct
86 Filled Polymers

A4.3  Medalia’s Aggregate Morphology Approach


Projected area A of one aggregate
[Source: A.I. Medalia, J. Colloid Interf. Sci., 24, 393–404, 1967]:

d: diameter of primary particle d

d2 Np: number of primary particle


A = π ⋅ ⋅ N pε
4 ε: empirical exponent from floc
simulation

   em: in Medalia’s
R
A
1 paper, ε = 0.87, but using
 4⋅ A  ε
his data from floc simulation,
= > NP = 
 π ⋅ d 2  in which d is set equal to
10 (nm), one finds in fact ε = 0.8471
Equivalent sphere model for a single aggregate
[Source: A.I. Medalia, J. Colloid Interf. Sci., 32, 115–131, 1970]

 VES: volume of d
3 equivalent sphere
π ⋅ D3 π  4 ⋅ A  2 D: diameter of
Ves = = ⋅
6 6  π  equivalent sphere D
A: projected area of
aggregate
3
4
Ves = ⋅ A2 A
3⋅ π

Volume VS of solid (carbon) within an aggregate:

Vs = Np⋅Vp VP: volume of one primary particle

π ⋅ d3
Vp = Ap: projected area of one primary particle
6
1
π ⋅ d3  4 ⋅ A  ε π ⋅ d3 π ⋅ d2
=> Vs = N p ⋅ = ⋅ Ap =
6  π ⋅ d 2  6 4
Typical Fillers for Polymers 87

Medalia’s relationship between void ratio [ratio Ves/Vs] and DPB absorption:
Ves 1 + vf
ec = CF ⋅ ⋅ .g – 1 = DBPA × ρ ⋅ 0.0115  CF = 0.765: correction factor
Vs C accounting for difference
between the projected area
of the equivalent sphere
and the projected area of
the aggregate (around 8.5%
reduction in D)
 vf = 0.46: void fraction for
randomly packed spheres
C = 1.4: correction for partial
fusion of primary particles
V C in aggregate
=> es = (1 + DBPA × ρ ⋅ 0.0115) ⋅
Vs CF ⋅ g ⋅ (1 + vf) g = 0.94: anisometry
correction factor for
non-perfect aligment of
aggregate’s main axis with
projection plan
ρ = filler specific gravity

[carbon black: ρ = 1.86 g/cm3]
C 1.4
= = 1.333
CF ⋅ g ⋅ (1 + vf) 0.765 ⋅ 0.94 ⋅ (1 + 0.46) DPBA = di-butylphtalate
absorption (cm3/100 g filler)
0.0115: correction for DBPA
end point (<0.85% of actual
void volume)
One has thus:
 4 3

Volume of equivalent sphere : Ves =  ⋅ A 2      4 3

 3⋅ π  Ves =  ⋅ A2 
 3⋅ π 
=>
 1

1  4 ⋅ A  ε π ⋅ d3 
Vs  ⋅
Volume Vs of solid (carbon): Vs =  4 ⋅ A  ⋅ π ⋅ d 
ε 3
 π ⋅ d 2  6 
 π⋅d 
2 6  

 3
d2
The projected area is: A = π ⋅ ⋅ N p =>
Ves
=
1
⋅ 4⋅
( π ⋅ d 2 ⋅ N pε ) 2 
4 Vs  3
 ε  ε1  3 
( N p ) ⋅ d 
 2 
[2 ⋅ π ]
 
88 Filled Polymers

 −1   ε 
Ves 1 ε 2 ε  21  Ves
= ⋅ N p ⋅ ( d ⋅ N p ) ⋅ ( N pε ) ε 
 ε + − 1
=> = N p 2 
Vs d VS

  ε
Ves ε + −1 C
Therefore: = N  2 
= ( 1 + DBPA × ρ ⋅ 0.0115 ) ⋅
CF ⋅ g ⋅ ( 1 + vf )
p
VS

 ε 
= (1 + DBPA × ρ ⋅ 0.0115) .1.333
 ε + 2 − 1
=> N p
2
N p = ((1+DBPA × ρ ⋅ 0.0115) .1.333) 3
⋅ ε−2

1
with ε = 0.87 N p = (( 1 + DBPA × ρ ⋅ 0.0115) .1.333) 0.305 N p = (( 1 + DBPA × ρ ⋅ 0.0115) .1.333)
3.279
  
1
with ε = 0.847 N p = (( 1 + DBPA × ρ ⋅ 0.0115) ⋅ 1.333) 0.27 N p = (( 1 + DBPA × ρ ⋅ 0.0115) .1.333)
3.697
  

ρ = 1.86

DBPA: = 30,32..160 [i.e., from N990:38cm 3 /100g to N356:157 cm 3 / 100g]


1 1
N p⋅Med (DBPA): = (1.333 ⋅ (1 + DBPA ⋅ ρ ⋅ 0.0115)) 0.305    N p (DBPA): = (1.333 ⋅ (1 + DBPA ⋅ ρ ⋅ 0.0115)) 0.27

1
<= explicit from A.I. Medalia, L.W.
N p⋅Med2 (DBPA): = (1.25 ⋅ (1 + 0.0115 ⋅ ρ ⋅ DBPA)) 0.305
Richards, J. Colloid Interf. Sci., 40,
233–252, 1972; g correction was not
applied (i.e. g = 1)

800
700
600 N330 : DBPA = 102cm3/100g
500 Np.Med(DBPA)
400
Np(DBPA)
Np.Med(102) = 114.124
300
Np.Med2(DBPA) Np(102) = 210.899
200
100
Np.Med2(102) = 92.435
0
0 90 180
DBPA
Typical Fillers for Polymers 89

A4.4  Carbon Black: Number of Particles/Aggregate


(Medalia vs. fractal approaches)
i = 0…17
N2 DPB Aggreg. Elem.Part.
absorpt. absorpt. diam. diam,
m 2 /g cm 3 /100g nm nm Data:
137.97 114 68.3 17.8  ρ: = 1.86 : carbon black specific gravity,
 N1110   
   126.00 133 76.7 18.8  g/cm3
 N121   116.75 115 77.3 20.7  ε: = 0.847 : scaling exponent
 N220   
   123.26 124 80.1 19.7 
 N234 
 N229 
 106.50 123 92.5 24.3 
 
   82.57 76 86.8 27.0 
 N326   
   79.45 102 105.0 30.1  F: = 2.44 : (mean) mass fractal dimension
 N330   92.10 120 103.0 25.8 
 N339    α: = 11 : front factor
   71.43 121 129.0 31.4 
 N351  Data: = 
 N358  88.05 157 136.0 30.22 
 
   84.30 155 136.0 30.2 
 N5500  
   41.12 117 234.0 56.8 
 N630   34.00 78 220.0 58.3 
 N650 
 
  DBPi : = (Data 1 ) i : DBP absorption,
 35.75 124 271.0 60.9 
 N660  cm3/100g
 N762 
 34.85 94 252.0 67.1 
 
   28.47 66 255.0 102.0  Di : = (Data 2 )i : mean aggregate
 N774    diameter, nm
   29.64 74 228.0 87.2 
 N990 
 9.09 38 483.0 291.0  di : = (Data 3 )i : mean particle dia
meter, nm

  2 

  1.15   3⋅ε − 2  
N1pi : = round  1.333.  1 + DBPi ⋅ ρ ⋅ , 0 : Number of particlees/aggregate
  100   
  according to Medalia

  D F 
N2 pi : = round α ⋅  i  ,0  : Number of particles/aggregate
  di   from fractal descripttion

x : 0, 100..600 800

 0  600
x: =  
 500 
400 N1p
x
200

0
0 200 400 600
N2 p, x
5
Polymers and Carbon Black

5.1  Elastomers and Carbon Black (CB)


5.1.1 Generalities
Whatever the filler, there are at least six families of parameters to consider
when studying how the properties of a given polymer are modified:

• The nature of the polymer, its chemistry, its macromolecular and


structural characteristics
• The chemical nature of the filler and its surface chemistry
• The average size and size distribution of particles
• The geometry and structure of particles
• The polymer/filler concentration ratio (or the filler volume fraction)
• The effects of the other ingredients in the compounds

No overall, coherent understanding of all those effects exists yet and


only several, partial aspects are documented, in various extent, for certain
classes of fillers. Due to its major role in rubber technology, CB is surely
the material that has so far captured most of the attention and therefore,
deserves a specific interest, whilst certain conclusions drawn in studying
polymer-CB composites are not necessarily applicable to other systems.
In most of their applications, elastomers are used in compounds where
CB plays the key role in reinforcing mechanical properties. Not all proper-
ties are improved when adding CB particles to an elastomer and a balance
of properties has to be achieved with respect to the application. In rubber
compounds, CB amounts to 20–30% volume fraction.
In rubber technology, a compounding formulation is generally described
in phr (part per one hundred rubber) of component. Let us consider a
typical general-purpose formulation for styrene-butadiene rubber (SBR)
(Table 5.1).

91
92 Filled Polymers

Table 5.1
Typical Carbon Black Filled SBR 1500 Compound
Ingredient phr Specific Gravity (g/cm3) Volume Fraction
SBR 1500 100 0.90 0.736
N330 carbon black 50 1.80 0.184
Zinc oxide* 5 5.57 0.006
Stearic acid 3 0.92 0.036
Processing oil 5 0.98 0.020
Antiheat 2 1.08 0.012
Antioxidant 1 1.17 0.006
Total 166 1.10 1.000
* Zinc oxide is nowadays often used as an 80% weight predispersion in EVA; specific gravity
of such a predispersion is 2.90 g/cm3, so 5 phr ZnO correspond to 6.25 phr predispersion.

The specific gravity of the compound is calculated as follows:

∑w i

ρcpd = i
(5.1)
∑ρ
w i

i i

where wi and ρi are, respectively the quantity (phr) and the specific gravity
of the ith ingredient.
The volume fraction of any ingredient i is calculated using:

wi × ρcpd
Φi = (5.2)
wcpd × ρi

where wcpd is the overall quantity (phr) of all ingredients in the formulation
(for instance 166 phr for the formulation described in Table 5.1).
It is fairly obvious that macroscopic properties of any complex polymer
system depend on the degree of dispersion of ingredients, in other words of
their spatial distribution within the volume of the material. This stresses the
importance of the preparation process, i.e., mixing operations that require
adequate equipments and procedures.
With CB, the reinforcing element is the aggregate. Therefore, optimum
mixing is achieved when all the filler is well dispersed at this level; ­ideally
no more agglomerates are present and breaking down aggregates into ele-
mentary particles, if arising, has been limited.
Basic considerations on optimum dispersion allow estimating the most
likely distance between aggregates. Let us indeed consider the most com-
pact spatial disposition of spheres of equal diameter; it corresponds to the
face centered cubic lattice (Figure 5.1). One may consider that an aggregate
occupies a spherical volume whose diameter is half the so-called Stokes
Polymers and Carbon Black 93

ε
da

Figure 5.1
Estimating inter-aggregates distances through a face-centered cubic lattice model.

diameter* da. Two typical distances between spheres (or aggregates) can
then be calculated with respect to filler volume fraction Φ, i.e.:
the longest distance:
 
 2 
ε= − 1  da
 3 12 Φ 
 π 
the shortest distance:

 
 2 
δ= − 1  da
 3 3Φ 2 
 π 

Table 5.2 gives results for a series of CB grades, with respect to volume
fractions 0.10 and 0.20, i.e., below and above the so-called ­percolation level
of around 0.13. As we will see later, significant changes occur in a number
of macroscopic properties, when filler content is above this level. In fact,
most practical rubber compounds have filler content significantly above the
percolation level (for instance Φ = 0.184 for the 50 phr N330 SBR compound
described in Table 5.1).

* The Stokes diameter is the volume of gyration of an anisometrical particle sedimenting in a


fluid, as measured using laser techniques.
94 Filled Polymers

Table 5.2
Estimating Inter-Aggregates Distances with a Face-Centered Cubic Lattice Model
Volume Fraction = 0.10 Volume Fraction = 0.20
Equivalent
Carbon Stokes Shortest Longest Average Shortest Longest Average
Black Diameter Distance Distance Distance Distance Distance Distance
Grade  (nm) (nm) (nm) (nm) (nm) (nm) (nm)
N110 85 40.3 74.6 57.5 23.2 50.5 36.9
N220 102 48.4 89.6 69.0 27.9 60.6 44.2
N234 97 46.0 85.2 65.6 26.5 57.6 42.1
N299 100 47.5 87.8 67.6 27.4 59.4 43.4
N326 106 50.3 93.1 71.7 29.0 63.0 46.0
N330 131 62.2 115.0 88.6 35.8 77.8 56.8
N339 104 49.4 91.3 70.3 28.4 61.8 45.1
N351 126 59.8 110.7 85.2 34.5 74.8 54.6
N356 138 65.5 121.2 93.3 37.7 82.0 59.9
N550 240 113.9 210.8 162.3 65.6 142.5 104.1
N660 251 119.1 220.4 169.8 68.7 149.1 108.9
N683 266 126.2 233.6 179.9 72.8 158.0 115.4
N762 275 130.5 241.5 186.0 75.2 163.3 119.3
N774 261 123.9 229.2 176.5 71.4 155.0 113.2
N990 436 206.9 382.9 294.9 119.2 258.9 189.1

Table 5.3
Interaggregates Distances in Compounds Developed for Optimal Reinforcement
SBR 1500 SBR1712
Carbon
Black Optimum Volume Interaggregates Optimum Interaggregates
Grade Fraction Distance (nm) Volume Fraction Distance (nm)
N110 0.201 21.4 0.207 20.2
N220 0.206 23.2 0.211 22.0
N234 – – 0.204 19.7
N326 0.250 24.0 0.253 23.4
N330 – 0.230 24.3
N339 0.216 20.6 0.215 20.3
N351 0.230 21.8 0.228 22.3
N375 – – 0.216 21.9

If one considers reinforcing blacks, for instance series N110–N356, with


a volume fraction of 0.20 (around 55 phr filler), typical shortest distance δ
­varies from 23 to 38 nm, and ε from 50 to 82 nm. As shown in Table 5.3, such
data are comparable with measurements made by Patel and Byers1 on ­several
compounds whose CB level was selected for optimal ­reinforcement, i.e., the
best balance of properties with respect to tire applications.
Polymers and Carbon Black 95

This simple exercise demonstrates that in most well dispersed CB filled


rubber compounds, average inter-aggregates distances are close to the aver-
age quadratic diameter of the polymer random coil, of the order of 50 nm for
diene elastomers with molecular weight of around 400,000 g/mole. In other
terms, in most rubber compounds, certain chains of elastomer are in contact
with at least two aggregates, whilst others are not.

5.1.2 Effects of Carbon Black on Rheological Properties


Generally the (shear) viscosity of a filled polymer increases with filler content,
as easily understood through simple hydrodynamic considerations. However
the flow behavior of filled polymers remains relatively simple only at very
low shear rate, with low structure materials of large particle size, for instance
glass micro-spheres, or thermal blacks with particle size larger than 500 nm.
Typical effects of carbon blacks on the shear viscosity function are illus-
trated in Figure 5.2, drawn using data published by Montes et al.2 Three types

Natural rubber/carbon black compounds at 100°C

Effect of filler level Effect of filler structure

1012 1012

20 phr N326 N326(20 phr)


1010 1010
Shear viscosity, kPa.s
Shear viscosity, kPa.s

108 30 phr N326 108 N110(20 phr)


10 Phr N326 N990(20 phr)

106 106
Gum Gum
104 104

102 102

100 100
102 103 104 105 106 102 103 104 105 106
Shear stress, Pa Shear stress, Pa

Low rate range Medium rate range High rate


range

M Capillary
“Sandwich” rheometer Mooney viscometer rheometer

Figure 5.2
Effects of carbon black on the shear viscosity function. (Data from S. Montes, J.L. White, N.
Nakajima, J. Non-Newtonian Fluid Mech., 28, 183–212, 1988.)
96 Filled Polymers

of viscometers were necessary to generate the data, all applying simple shear
flow. The low rate range was investigated with a (prototype) “sandwich” rhe-
ometer, the medium range with a variable speed Mooney viscometer and the
high rate range with a capillary rheometer. One notes incidentally that the
viscosity is plotted vs. the shear stress.
The effect of filler level is shown in the left graph. At 100°C, gum natural
rubber (NR) exhibits the expected behavior of a pseudo-Newtonian plateau
at low shear stress (and hence low shear rate), then a shear thinning behavior.
As filler content increases, the Newtonian plateau progressively disappears
and above 20 phr, the viscosity variation with decreasing shear stress is such
that a yield stress behavior is suggested.
The effect of the structure is shown in the right graph. As can be seen,
as the filler structure is more and more complex (from N990 to N326, then
N110), the Newtonian plateau disappears, and again a yield stress behavior
is suggested for the reinforcing carbon blacks.
There is thus, a striking parallelism between the effect of filler content
at constant structure, and the effect of filler structure at constant level. We
note that the η = F(σ) representation of the viscosity function is important
in observing the above effects, particularly the occurrence of a yield stress.
Using the (more traditional) plot of η vs. the shear rate ( γ ), only the pro-
gressive disappearance of the Newtonian plateau would have been detected,
with maybe a slight concave curvature appearing at very large filler level, as
illustrated in Figure 5.3.
Much has been published about the yield stress behavior of CB filled
compounds but it is worth underlining here the experimental difficulties
in assessing the shear flow behavior of highly viscous materials in the very
low shear rate (or stress) range. No commercial equipment exists for vis-
cosity measurement below γ = 10−2 s-1, and as shown in Figure 5.2, Montes
et al. used a prototype plan-plan rheometer, developed at the Polymer
Engineering Institute, University of Akron, OH and operated through a
creep technique. Drag flow was obtained by moving a plan through the
action of a dead weight; a cathetometer was used to measure displacement
over various time intervals. It is obvious that cathetometer accuracy reading
and operator patience were key elements of the experiment. Any yield stress
value that can be derived from such experiments is an extrapolated value; so
strictly speaking, one does not measure it, as quoted in the controversial lit-
erature on the subject.* Recently Barrès and Leblanc3 developed a prototype
­sliding cylinder rheometer that uses an optical transducer with the capabil-
ity to detect displacements in the micron range, and a computer to record

* Note that there are many (complex) fluids that do not exhibit a Newtonian plateau at low
shear rate (stress) and whose shear viscosity function feeds the controversy on the existence
of a yield stress. As noted by Barnes, when the flow is so slow than ages are necessary to
detect it, at least one could consider that the yield stress is an engineering reality (H. Barnes.
The yield stress—a review or “παντα ρει” —everything flows? J. Non-Newtonian Fluid Mech.,
81, 133–178, 1999.)
103
SBR 1500 – N220 carbon black
T = 100°C

102
: Gum SBR

Shear stress, kPa


Flow curve σ = f(γ) : Φblack = 0.10
Polymers and Carbon Black

: Φblack = 0.20
101
10–2 10–1 100 101 102 103 5×103
Shear rate, s–1
104 104
Shear viscosity

103 function η = f(γ) 103

102 102

101 101
Shear viscosity

Shear viscosity, kPa.s


Shear viscosity, kPa.s
100 100 function η = f(σ)

10–1 10–1
10–2 10–1 100 101 102 103 5×103 2×101 102 103
–1
Shear rate, s Shear stress, kPa

Figure 5.3
Shear viscosity plots vs. carbon black level.
97
98 Filled Polymers

High cis-1,4 polybutadiene/N330 (50 phr) - T = 100°C


106 106
105 η = f(σ)
105

Shear viscosity, kPa.s


104
104 103
Shear viscosity, kPa.s

102
103 •
v 101
2
10 100
M
10–1
101
10–2
100 104 105 106
Shear stress, Pa
10–1 SCR creep mode
SCR constant rate
10–2 Mooney viscometer
10–5 10–4 10–3 10–2 10–1 100 101 102 103 104 Capillary rheometer
Shear rate, s–1
Figure 5.4
Shear viscosity function of a carbon black filled BR compound.

data over very long periods of time (up to 24 h, in certain cases). Shear vis-
cosity data obtained with this instrument compare well with (extrapolated)
ones generated with both the Mooney viscometer and a capillary rheometer.
As shown in Figure 5.4 for a CB filled BR compound, they did not detect
any Newtonian plateau down to 10−5 s−1, but the η = F(σ) plot would hardly
be considered as suggesting the existence of a yield stress, at the least in the
shear rate range investigated.
As a matter of fact, there are a number of complex systems that, provid-
ing the appropriate equipment is used (for instance the so-called “controlled
stress” rheometers), are found to exhibit a drastic fall of the viscosity, by
several orders of magnitude, over a narrow range of shear stress. When
approaching this critical region from the high stress range, it seems thus that
the viscosity goes to infinity at a certain minimum stress.
Thanks to their relatively lower viscosity, food and cosmetic products allow
this type of behavior to be easily documented, when performing experiments
with controlled stress rheometers. There are commercial versions of such
instruments, essentially rotating systems (parallel disks or cone-and-plate),
which have the capability to measure extremely small rotation rates (in the
10−8 rad/s, i.e., one revolution in 20 years!). Experiments performed in such
conditions are called creep testing using controlled torques as low as 10−7 Nm
with a resolution of 10−9 Nm. Figure 5.5 shows an example of the shear viscos-
ity function (vs. shear stress) measured on a typical cosmetic product (body
cream), using such a controlled stress rheometer.
As can be seen, measurements made at stress higher than 50 Pa and extrap-
olated toward lower shear stress would allow detecting an (apparent) yield
stress. In fact, a drastic fall of viscosity (around two millions time) occurs at
Polymers and Carbon Black 99

Controlled stress rheometer σc = 44.6 Pa


107
η1 = 3.16×106 Pa.s
106
Drastic fall of
σ1 = 13.8 Pa
105 viscosity for a
slight variation of stress
Shear viscosity, Pa.s

104

× 2.000.000
103

102

101
Body cream σ2 = 79.5 Pa
(room temperature)
100
×2 η2 = 0.39 Pa.s
10–1
100 101 102 103
Shear stress, Pa
Figure 5.5
Typical shear viscosity function of a (complex) cosmetic material.

a critical stress of 44.6 Pa, splitting the flow behavior in two nearly constant
viscosity regions, one in the 106 Pa.s range, the other in the 10−1 Pa.s range.
Such a behavior might also exist with filled high molecular weight ­polymer
systems, for instance rubber compounds, but it has so far never been observed
(to the author’s knowledge). Consequently, even if the yield stress exists only
because the shear rate/shear stress range of observation does not allow the
above behavior to be observed, simple yield-stress-containing equations for
viscosity properties are useful, at least over a limited range of stress/rate.
A number of mathematical models have been proposed for yield stress flu-
ids, not all perfectly coherent however. If a shear thinning material is tested
with one of several rheometers and the results plotted in terms of shear stress
σ vs. shear rate γ (the so-called flow curve) using linear scales, one is nearly
bound to the conclusion that there is a yield stress σc and that the best man-
ner to model the observed behavior consists in considering the following
equality: σ = σ c + f ( γ ) .
As we will see below, nearly all proposed models follow this approach,
but with respect to the definition of the shear viscosity function, i.e.,
η = σ/γ = f ( γ ) = F(σ ), we consider that a criterion of coherence, if not validity,
of such models is that they allow equations to be derived, either as η = f ( γ )
or as η = F(σ).
Herschel and Bulkley4 combined the power law model with a yield stress σc:

σ = σ c + K γ n (5.3)
100 Filled Polymers

so that the following equations are obtained for the shear viscosity
function:
σc
η = f ( γ ) : η( γ ) = + K γ ( n−1) (5.3a)
γ
1/n
 K 
η = F(σ): η(σ ) =  σ (5.3b)
 σ − σ c 

As illustrated in Figure 5.6, the Herschel–Bulkley equation corresponds to


viscosity functions without Newtonian region, which exhibit a significant
upward curvature as the shear rate/shear stress is decreasing. The η = F(σ)
shows indeed a viscosity that goes to infinity as the stress decreases toward
a critical value. It is fairly obvious that, when fitting shear viscosity data with
an equation that explicitly states that there is a yield stress (Equation 5.3,
for instance), a value for σc will be obtained, which has very limited mean-
ing, if any, when the investigated shear stress range is excessively far from
where the fit parameter is obtained. In the high rate/stress region, the shear

• Herschel–Bulkley equation
Flow curve σ = f (γ)
15
K = 0.9
n = 0.4 σ = σc+ Kγn
10
Shear stress


σc
η(γ) = •
+ K γ• (n–1)
γ
5
1
σc = 0.5 K n
η(σ) = (
σ – σc ) σ
0
0 200 400
Shear rate


Viscosity function η= f (γ) Viscosity function η = f (σ)
102
105
σc = 0.50
101
103
Shear viscosity

σc = 0.50
Shear viscosity

100
σc = 0.25 101
10–1

10–2 10–1 σc = 0.25

10–3 –1 10–3 –1
10 100 101 102 103 10 100 101 102
Shear rate Shear stress
Figure 5.6
Herschel–Bulkley model for yield stress fluids.
Polymers and Carbon Black 101

thinning behavior met by the Herschel–Bulkley equation corresponds to the


power law model, not always suiting the downward curvature of the viscos-
ity function observed in practice with complex polymer systems.
The so-called Bingham fluid obeys to the following equation:

σ = σ c + η ⋅ γ (5.4)

In other terms, above a critical shear stress, it flows as a Newtonian fluid


of (constant) viscosity η. It follows that a fluid obeying the Herschel–Bulkley
model is sometimes called a generalized Bingham fluid, since with n = 1 and
K = η in Equation 5.3, one obviously obtains Equation 5.4. The three fit param-
eters of the Herschel–Bulkley equation can be reduced to two, when consid-
ering that n = 0.5. This was in fact the approach used by Casson5 in proposing
the following model:

σ = σ c + K γ   or  σ = σ 0c .5 + 2 ⋅ σ 0c .5 ⋅ K ⋅ γ 0.5 + K 2 ⋅ γ (5.5)

Explicit formulas for both shear viscosity functions are easily derived, as
follows:
σ σc
η = f ( γ ) : η( γ ) = c + 2 K + K2 (5.5a)
γ γ
2
 K 
η = F(σ): η(σ ) =  σ 2 + σ × σ c + 2 σ σ − σ c   (5.5b)
 σ c − σ 

As illustrated in Figure 5.7, the Casson model corresponds indeed to a


η( γ ) function without Newtonian plateau and an upward curvature in the
low shear rate region significantly depending on the magnitude of the yield
stress; but the shear thinning behavior does not correspond to observations
on filled polymer systems. Similar comments are made for the η(σ) function.
White et al.6 submitted a model (the White–Wang model) that was used
with some success in a number of works made at the Institute of Polymer
Engineering, Akron, OH. With respect to the flow curve, this model writes
as follows:
A γ
σ = σc + (5.6)
1 + B γ 1− n

where A and B are fit parameters whose ratio A/B corresponds to the param-
eter K of a power law σ = K γ n used to model the high shear rate range, and
n is the flow index. An explicit equation is obtained for the viscosity function
η( γ ), i.e.:
σc A
η = f ( γ ) : η( γ ) = + (5.6a)
γ 1 + B γ 1− n
102 Filled Polymers


Flow curve σ = f (γ)
40 Casson equation
K = 0.2 √σ = √σc + K√γ

σ = √σc + 2√σc K√γ +K 2γ


• •

η (γ) = σ + 2√σ K + K 2

Shear stress

c c
20 •
γ

γ
2
η(σ)= (σ2+ σ. σc + 2 σ√σ– σc)
(σ K– σ )
c
σc= 5
0
0 100 200
Shear rate
Viscosity function η= f (γ)

Viscosity function η= f (σ)
101 102
σc= 5
101
Shear viscosity

σc= 5
Shear viscosity

100

σc= 1 100
10–1
10–1
σc= 1

10–2 10–2
100 101 102 103 100 101 102 103
Shear rate Shear stress

Figure 5.7
Casson model for yield stress fluids.

but there is no single solution for η(σ) Indeed, by substituting γ = σ/η in


Equation 5.6, one obtains:

A σn
η = F(σ): η(σ ) σ n−1 + η(σ )n B = (5.6b)
(σ − σ c )

Figure 5.8 illustrates the capabilities of Equations 5.6 and 5.6a. Despite its
limitations, the White–Wang model suits well certain observations made on
CB filled rubber compounds, particularly the behavior of the shear viscos-
ity function η( γ ) in the low shear rate region, as reported in a number of
publications.
Even if the White–Wang model remains open to discussion, the yield stress
data that can be derived from experimental flow curves by fitting Equation
5.6 are likely to be similar to values that would be obtained with other mod-
els, for instance the Herschel–Bulkley equation. It is consequently interest-
ing to pay attention to yield stress data that were obtained at the Polymer
Engineering Institute, using this model.
Polymers and Carbon Black 103

Flow curve σ = f (γ• )


15 White–Wang equation
A=2
B = 0.8 Aγ•
n = 0.3 σ = σc+
10
Shear stress

1 + B γ• (1–n)

σc A
5 η(γ• ) = +
γ• 1 + B γ• (1–n)
σc = 0.5

0
0 50 100
Shear rate

Viscosity function η= f (γ• ) Viscosity function η = f (σ)


102

σc= 0.50
101 No single solution for η = f (σ)
Shear viscosity

100 σc= 0.25


η(σ) . σ n–1 + η(σ)n.B = A . σ n
(σ – σc )
10–1

10–2
10–2 10–1 100 101 102
Shear rate
Figure 5.8
White–Wang model for carbon black filled compounds.

Table 5.4 is a compilation of yield stress data for various rubber compounds.
As previously described (see Figure 5.2) three rheometrical techniques were
used to obtain the shear viscosity data, and the yield stress values were
derived by fitting results with Equation 5.6a. Rubber nature and CB type
are indicated, as well as the volume fraction ΦBlack and the crushed di-butyl
phthalate (DBP) absorption numbers (cDBPA) of CB. As can be seen, both the
filler level and the reinforcing character of CB do affect σc.
One might consider that the yield stress behavior of CB filled compounds
does reflect the response to (shear) stress of some kind of tri-dimensional
structure, either occurring through contacts between aggregates or, most
probably, resulting from rubber–filler interactions and the associated com-
plex morphology. We will elaborate further this comment in subsequent sec-
tions, but would such considerations be valid, one should expect the yield
stress to be commensurate with a complex parameter in which the volume
fraction and an assessment of the aggregate structure are involved. A very
simple manner to probe this argument consists in plotting yield stress data
vs. the corresponding product ΦBlack × cDPBA. As shown in Figure 5.9, a loose
relationship is indeed observed.
104

Table 5.4
Yield Stress Data on Various Filled Rubber Compounds
Carbon Black Volumic Fraction Yield Stress
Elastomer Type Φ cDPBA (dm3/kg) Φx cDPA (kPa) Source
Natural rubber (RSS1) N326 0.2 0.67 0.134 60 116
Natural rubber (SMR L) N220 N550 N762 0.2 0.2 0.2 0.99 0.84 0.59 0.198 0.168 0.118 119 79 62 118
Natural rubber (SMR 5) N110 N326 N326 0.2 0.2 0.3 1.01 0.67 0.67 0.202 0.134 0.201 140 44 170 2
Polyisoprene Cis-1,4 (Shell IR305) N326 0.2 0.67 0.134 38 116

Polyisoprene Cis-1,4 (Natsyn 2200) N326 0.2 0.67 0.134 42 116


SBR1500 N326 N326 N326 0.1 0.2 0.3 0.67 0.67 0.67 0.067 0.134 0.201 6 13 100 6, 117
N330 N330 0.2 0.3 0.87 0.87 0.174 0.261 63 138
SBR Emulsion Goodyear N220 N550 N762 0.2 0.2 0.2 0.99 0.84 0.59 0.198 0.168 0.118 133 83 63 119
Plioflex 1500
SBR Sol (23.5 Styrène) N220 N550 N762 0.2 0.2 0.2 0.99 0.84 0.59 0.198 0.168 0.118 130 82 62 118, 2
Firestone Duradene 706 N326 0.2 0.67 0.134 100
EPDM N330 0.2 0.87 0.174 58
Chloroprene rubber CR N330 0.2 0.87 0.174 56
Filled Polymers
Polymers and Carbon Black 105

180
160
Apparent yield stress, kPa 140
120
100
80
60
40
20
0
0 0.1 0.2 0.3
Volume fraction × cDBPA
Figure 5.9
Relationship between yield stress (from flow curves at 100°C), carbon black structure, and level
in filled rubber compounds.

As we have already underlined, adding CB to an elastomer gives a


compound that has a strong nonlinear viscoleastic character, appearing
amongst other effects, by the quasi-suppression of the linear region in
the shear viscosity function (which no longer exhibits a Newtonian pla-
teau at low shear rate or stress). It follows that other typically nonlinear
effects are exacerbated by the presence of CB particles and, as can be
expected, the higher the reinforcing character of the filler, the stronger
those effects.
Several typical examples allow this aspect to be illustrated. Let us ­consider
the so-called stress overshoot effect, i.e., the peak stress observed in stress
growth experiments when the suddenly applied shear rate is very high.
This effect is typically non linear and has been well documented through
­laboratory experiments on a number of simple polymer systems.7 With
filled rubber materials, experimental data on stress overshoot are scarce,
likely due to ­difficulties in handling such very stiff materials in conven-
tional ­rheometers (i.e., parallel disks or cone-and-plate). Experiments by
Montes et al.2 are therefore, worth considering. These authors used a vari-
able speed Mooney viscometer to perform stress growth experiments on
NR and SBR 1500 compounds with various types and levels of CB. Figure
5.10 describes their experimental approach and gives typical results on
NR compounds, expressed as ratio of the “peak” viscosity over the stabi-
lized viscosity.
The stress overshoot is amplified by the presence of CB, and the more rein-
forcing the filler, the stronger the effect. Such experiments, whilst obtained
in laboratory conditions, provide results that correspond very well (at least
qualitatively) with observations on factory floor. Rubber engineers are indeed
well aware of the difficulties associated with the starting-up of extrusion
lines when processing highly reinforced compounds. Machine temperature
profile and extruder screw speed must be carefully monitored right from the
106 Filled Polymers

Variable speed Mooney viscometer (Montes, White, Nakajima–1988)

Peak

Torque
Maximum
viscosity ηmax
Stabilized
viscosity η∞

Cst shear rate


1.6 Time
Natural rubber (SMR 5)
1.5
Carbon black : 20%
Temperature : 100°C N110
1.4
ηmax
η∞ 1.3
N326
1.2

1.1 N990

1.0
0.02 0.1 1 2
Shear rate, s–1
Figure 5.10
Stress overshoot experiments on carbon black filled compounds.

beginning of the operation in order to avoid excessive head pressure, and


quite long periods (while wasting material) are sometimes necessary before
the extrusion line is stabilized.
Other easily observed effects of CB on the elastic character of rubber
compounds are made when considering the so-called “extrudate swell”
(or postextrusion swelling). All other parameters constant, extrudate
swell decreases when the CB level increases, 8 as illustrated in Figure
5.11. From a technological point of view, this means that in rubber tech-
nology, postextrusion swelling is at least one order of magnitude below
what is commonly observed with pure polymers (for instance polyole-
fins). Consequently extrusion die design and postextrusion equipment are
notably different.
At constant CB level, the more complex the filler structure the lower the
extrudate swell9 as illustrated by Figure 5.12. An immediate practical con-
sequence of such effects is that the dimension stability of highly reinforced
tire tread bands is relatively easy to achieve on the factory floor, by control-
ling the longitudinal shrinkage associated with the swelling (normal to the
extrusion axis).
Such effects, obviously beneficial and exploited in rubber processing tech-
nology, remain however difficult to fully explain. Indeed, whilst extrudate
swell decreases with increasing CB volume fraction, the die entrance pressure
Polymers and Carbon Black 107

Natural rubber (SMR10)/N220 carbon black


Capillary rheometer at 100°C
60

10 phr
50 25 phr
(Relaxed) extrudate swell, %

40
40 phr

30

20

60 phr
10

0
0 0.1 0.2 0.3 0.4
Shear stress (MPa)
Figure 5.11
Effect of carbon black level on postextrusion swelling.

100 SBR 1500


90 Carbon black (50 phr)
Extrudate swell, %

80
70
60
50
40
30
0.2 0.4 0.6 0.8 1 1.2 1.4
DBP absorption (cm3/g)
Increasing complexity of filler particle structure

Figure 5.12
Effect of carbon black structure on postextrusion swelling.

loss significantly increases with higher filler content. Such a behavior is totally
different from what is observed on compounds made with CB and saturated
polymers. For instance, Robertson et al.10 reported experiments on the non-
linear rheology of hyperbranched polyisobutylene, notably capillary rhe-
ometry at 205°C on linear and branched polyisobutylene compounded with
50 phr N339 and 5 phr processing oil. Figure 5.13 is drawn using extrudate
108 Filled Polymers

1.50
Capillary flow at 205°C; 1.5 mm dia. dies
PIB cpds (50 phr N339; 5 phr oil)
Extrudate swell 1.25

1.00

0.75 Linear polyisobutylene


Branched polyisobutylene
0.50
0 2 4 6 8
Entrance pressure, MPa
Figure 5.13
Postextrusion swelling vs. entrance pressure drop, as observed on compounds of carbon black
with saturated polymer melts.

swell data (measured after nine days resting at room temperature) and the
­corresponding entrance pressure drop data Pent (assessed through the Bagley
method). Despite a large scatter on Pent, it appears that there is direct propor-
tionality between postextrusion swelling and entrance pressure drop. One
notes incidentally that the structure of the polymer has no significant effect
on the relationship observed.
Robertson et al. give no information regarding likely interactions between
CB and polyisobutylene and one might presume that their observations
reflect only hydrodynamic effects. Either linear or branched, polyisobuty-
lenes are polymers with a high chain flexibility, like any elastomer, but unlike
diene rubbers, there is no chain unsaturation that could, on one hand signifi-
cantly increases the mean diameter of the random coil (and hence increases
the probability for one macromolecule to be in contact with at least two CB
aggregates), and on the other hand favors interaction of topological nature
between polymer segments and the surface of filler particles, as we shall dis-
cuss below.

5.1.3  Concept of Bound Rubber (BdR)


When an elastomer and a reinforcing filler are mixed, strong interactions
occur in such a manner that, at room temperature, a good solvent of the
­polymer can extract only a free rubber portion, leaving a highly swollen
rubber–filler gel. BdR is, by definition, the rubber content of that gel, and
the most obvious proof that a heterogeneous, complex structure has been
formed during the mixing operations. Known for more than 80 years in the
case of CB filled compounds, it is considered one of the major factors in rein-
forcement and often a global measure of surface activity of the filler, because
of the striking parallelism between the amount of BdR and the reinforcing
capabilities of the black. Rubber–filler ­interactions readily occur in the early
stages of the mixing process and there are consequently direct relationships
Polymers and Carbon Black 109

Table 5.5
Effect of Carbon Black Size on Bound Rubber; Literature Data
Bound Rubber (%)
Carbon Black
(50 phr) SBR (Source) BR (Source)
N110 42.2 ± 0.8 (12,111,112) 30.0 (113)
N220 36.0 ± 1.2 (12,49,111–114) 27.0 ± 2.1 (113,114)
N330 32.1 ± 1.4 (12,111,112) 27.1 ± 4.6 (113,114)
N339 36.1 ± 2.9 (12,111,112) –
N347 34.7 ± 1.0 (111,112) 27.0 (113)
N351 32.6 ± 3.2 (12,49,111–114) –
N375 36.8 ± 3.4 (12,49) 29.1 (113)
N440 22.9 ± 0.4 (12,111) –
N550 – 20.1 (113)
N539 24.4 ± 1.1 (12,111) –
N568 25.2 ± 0.1 (12,111) –
N650 19.9c
N660 – 12.9 (114)
N765 21.5 ± 0.1 (12,111) –
N774 – 11.0 (114)

between BdR and rheological properties of rubber compounds, as outlined


in a recent review paper.11
For compounds made with a given elastomer and at constant filler level
(with all other compounding ingredients constant), BdR increases with
the reinforcing character of CB. For instance, BdR increases with decreas-
ing elementary particle size, and increases as the structure is more and
more complex, as clearly seen when compiling literature data (Table 5.5
and Figure 5.14), despite some differences between authors. In Table 5.5,
the particle size effect is readily seen when comparing how BdR evolves
in BR compounds with 50 phr of CB of the N110, N220, N330, N550, and
N660 grades. Within the same family, for instance the N300 series, increas-
ing aggregate complexity is considered through N330, N339 and N347 in
SBR compounds.
Figure 5.14 compares data reported by different authors12,13 on SBR 1500
compounds filled with 50 phr CB. As can be seen, the effect of specific area of
CB (which reflects the complexity of aggregates) on BdR is quite consistent:
the more reinforcing the CB, the higher the BdR.
Providing the filler content is above or near a critical level of 12–13%, BdR
measurement yields a swollen, coherent sample at the end of the extraction
process. This suggests that there are rubber–filler ­interactions, strong enough
to resist the solvatation process, which consequently lead to a tri-dimen-
sional morphology in uncured filled rubber compounds. Microphotography
evidences of such a 3D structure were published in the early 1970s14,15 and
110 Filled Polymers

45 N110
SBR 1500 compounds N375
40 50 phr carbon black N339
N347 N220
35 N351 N285 N219
N330 N121
30 N327 N234
Bound rubber, %

N568 N339 N375 N242 N110


25 N539 N326t N220
N440 N330
N765
20 N539
N683 N326n
N550
15 N765 N660
10 N762
Dannenberg (1986)
5
Wolff, Wang, Tan (1993)
0
0 20 40 60 80 100 120 140
Specific area, m/g2
Figure 5.14
Effect of carbon black structure on bound rubber; literature data.

several authors used advanced nuclear magnetic resonance (NMR) tech-


niques to obtain detailed information about how the molecular flexibility
is modified by the presence of filler particles. Three distinct regions in the
filled rubber, with different degrees of molecular mobility exhibited by the
constituent rubber chains are generally reported, irrespective of the chemi-
cal nature of the rubber.16–18
In a well-dispersed, uncured compound, isolated aggregates induce
restrictions to mobility of rubber chains, in such a manner that three rubber
regions can be distinguished. Indeed NMR relaxation experiments reflect
the heterogeneity of filled rubber materials, with a fast decay in the 20 µs
range that is assigned to the immobilized part of the BdR shell. At longer
times (above 100 µs), there is mobile fraction that contributes to the relaxation
and in between (in the 50 µs range), there is an intermediate mobility com-
ponent. NMR relaxation curves are decomposed into components, using the
appropriate number of distribution functions.
For instance, Yatsuyanagi et al.19 performed pulse NMR measurements
at 100 MHz on a series of SBR 1502 compounds with various levels of HAF
(~N330 and ~N339) and SAF (~N212) and analyzed the results with respect
to BdR content (72 h extraction in toluene at room temperature). By con-
sidering solid echo sequence, the proton spin–spin relaxation time T2 was
obtained and reduced in several components by fitting the signal with
a set of Weibull functions. Above −30°C, filled compounds exhibit three
­components in T2, a long T2L, an intermediate T2M and a short one T2S that are
assigned to three different types of BdR. The unfilled compound shows a
simple signal that corresponds to T2L and, as illustrated in Figure 5.15, there
are clear relationships between T2M and T2S and BdR content. In agreement
with other publications, the T2 components can therefore be considered
Polymers and Carbon Black 111

500 40 phr : 30 60 90
N330

Bound rubber components, % gum


400 N339
T2L
N121
300 30
T2 components, µs

Extractable rubber
200
20
100
80
60 T2M 10
40 Loosely BdR
T2S
20
Tightly BdR
0 0
0 10 20 30 40 50 0 10 20 30 40 50
Bound rubber, % gum Bound rubber, % gum
Figure 5.15
Bound rubber and NMR results.

Loosely bound rubber


10–36% elastomer; Unbound (extractable) rubber
thickness : 3.0–6.6 nm 60–90% elastomer

Connecting
filament

Tightly bound rubber Filler particle


2–4% elastomer; Equiv. spherical diameter : 100 nm
thickness : 0.4–1.3 nm
Figure 5.16
Tridimensional representation of the morphology of carbon black filled rubber compounds.

as a clear indication of three types of rubber in a filled ­compound: the


­extractable (i.e., unbound) fraction and two types of BdR, with different
molecular mobility.
There are thus various experimental results and common sense observa-
tions that support the morphology of an uncured CB filled compound as
depicted in Figure 5.16.20
112 Filled Polymers

Very close to filler particles, there is a thin layer of tightly BdR, which is
likely to behave in a flow field exactly as the aggregate. Then there is a region
of loosely BdR, i.e., chains attached to the particles—through the tightly BdR
region—but able to undergo very large deformation during flow. This region
eventually forms connective filaments between rubber–filler aggregates. The
third portion is the unbound rubber, so-called because it can readily be extracted
from the uncured compound by a good solvent of the elastomer. Obviously
this rubber region does not interact with the filler particles. Extractable rub-
ber accounts for 70–90% of the gum elastomer of the compound, depend-
ing on the formulation. The tightly and loosely BdR fractions are up to 30%
of the gum. The connecting filaments, which are readily seen on published
microphotographs,14,15 ensure the coherence of the swollen rubber-black gel
in a good solvent.

5.1.4 Bound Rubber at the Origin of Singular Flow Properties


of Rubber Compounds
It is important to underline here that the schematic description of rubber–CB
systems given in Figure 5.16 has to be considered as an “instant” view of
materials, which remain essentially (pseudo) fluids above their glass temper-
ature, i.e., practically in their whole processing and application temperature
windows. We mean that rubber–CB interactions are likely to be dynamic,
through continuous adsorption–desorption processes, at equilibrium at any
given temperature. BdR content is therefore nothing else than the assess-
ment of an equilibrium state at the temperature of the test, and therefore, is
expected to decrease as extraction temperature increases (as readily observed
indeed). If it were not the case, such complex materials would not have the
capability to flow.
Materials which such a complex structure cannot of course exhibit a sim-
ple rheological behavior, and relatively easy arguments may be produced
to explain—so far qualitatively—how this complex structure affects most
of the flow singularities of rubber compound. As illustrated in Figure 5.17,
­typical nonlinear effects observed with CB filled compounds appear as logi-
cal consequences of such a soft three-dimensional network of complex rub-
ber–aggregate entities with connective filaments.
The suppression of the Newtonian plateau in the shear viscosity function
when the filler content is above 12–13% can be seen as the extension in the
low rate range of the nonlinear viscoelastic character, otherwise observed
on pure, homogeneous polymers in the high rate range. However such a
­nonlinearity is “internal” (i.e., morphology induced, or “intrinsic”) to filled
compounds, a part of their basic character, in contrast to the “external” (i.e.,
strain induced) nonlinear viscoelasticity of pure polymers that appear when
stress conditions overcome a certain limit. With respect to Figure 5.16, it is
easy to understand that as the applied stress or rate of strain reduce, the
system responds more and more as an elastic network and, therefore, the
Morphology of carbon black
filled rubber compounds Soft three-dimensional network
of bound rubber–filler particle
units with connective elements
Unbound rubber

Carbon black aggregate


Polymers and Carbon Black

Bound rubber

Essentially non linear


flow properties

Suppression of Stress dissipation effects Enhanced flow Enhanced Reduced chaoticity


Newtonian plateau (lower extrudate swell) anisotropy effects wall slip (smoother melt
(yield stress behavior) (converging flow) fracture)

Observed according to conditions and rate of deformation

Figure 5.17
Carbon black filled rubber compound morphology and nonlinear flow properties.
113
114 Filled Polymers

apparent viscosity tends to excessive values. Compounds with high levels


of reinforcing fillers exhibit practically no flow before stress values as high
as 105 Pa are reached, as a consequence of the relative amount of BdR, i.e.,
elastomer chains with restricted mobility, that is increasing with both the CB
content and/or the aggregate structure.
The decrease in extrudate swell magnitude with higher CB content, as well
as the lower severity of turbulent flow defects (the so-called extrudate melt
fracture) can be assigned to a dampening effect due to the filler, through
stress dissipation effects involving the complex rubber–­aggregate units and
their connective filaments. In their displacement (in the main flow direction),
units influence each other and either limit the storage of elastic energy by
increasing the viscous dissipation term, or favor the elastic release through
local microrelaxation processes.
The slippage at the wall of processing equipment is a common sense
­observation on factory floor, likely affected by minor compounding ingredi-
ents such as oil, plasticizers, processing aids, and other low molecular weight
chemicals. Local drag flow mechanisms of rubber–aggregate flow units, not
only throughout the bulk of the material, but also close to the wall, obviously
enhance this effect.
Flow anisotropy effects are important and easily observed in a number of
rubber processing operations and the so-called “grain effect” in calendering
is a well-known example. In injection molding, flow anisotropy effects have
been shown to persist after vulcanization, as illustrated in Figure 5.18. Such
effects arise from interactions between rubber–filler units in such a manner
that self-organization of the soft network occur in the main flow direction,
particularly when converging or diverging conditions prevail. In a sense the
rubber–filler soft network can be viewed as an “elasticity dissipation struc-
ture” since such an organization process (which by the way has an antien-
tropy character) consumes a part of the strain energy, otherwise stored in an
elastic (i.e., recoverable) manner.

5.1.5  Factors Affecting Bound Rubber


Qualitatively, literature data on BdR are quite coherent but, quantitatively,
a large scatter is observed, as illustrated by Table 5.6 with data on SBR 1500
compounds from different authors. For a given CB grade, data reported by
different authors are within a 5–6% bracket, but the expected effect of filler
grade is clearly seen.
Simple in its principle, BdR measurement is in fact affected by a number
of factors, as follows:

1. The exact nature and formulation of the compound (elastomer type,


its molecular weight and MWD, unsaturation level, chemical nature
and stability, filler type, its characteristics, level, and dispersion state;
effects of other mixing ingredients)
Polymers and Carbon Black 115

Tensile test on cured samples


100

300% modulus, kPa


Flow anisotropy in
injection moulding 80
60 Natural rubber compound
40 60 Shore A
20 Vulcanization at 160°C
Test sample cut 90°
0
vs. flow direction 0 45 90
Angle between flow and stretched directions, °
100

300% modulus, kPa


80 Chlorobutyl rubber compound
60 Shore A
60
Vulcanization at 160°C
40
180 mm

20
0
0 45 90
Angle between flow and stretched directions, °
100 EPDM compound
300% modulus, kPa

Test sample cut 0° 80 60 Shore A


vs. flow direction Vulcanization at 160°C
60
2 mm
40
20
0
0 45 90
Angle between flow and stretched directions, °
Figure 5.18
Flow induced anisotropy effects in filled rubber compounds. (Drawn using data from
K. Nakashima, H. Fukuta, M. Mineki, J. Appl. Polym. Sci., 17, 769–778, 1973.)

Table 5.6
Bound Rubber of 50 phr Carbon Black SBR 1500 Compounds; Literature Data
Bound Rubber (% Initial Gum)
Carbon Standard
Black Literature Data (Source) Average Deviation
N220 34.8 (115) 34.8 (112) 37.4 (12) 26.3 (13) 32.8 5.8
N330 29.7 (115) 30.5 (112) 32.2 (12) 22.9 (13) 28.5 4.9
N339 34.2 (115) 37.2 (112) 38.3 (12) 27.0 (13) 34.2 6.0
N539 – – 26.3 (12) 17.6 (13) 21.9 6.1
N765 – – 21.4 (12) 13.7 (13) 17.6 5.4

2. The compound preparation procedure (mixing parameters, energy,


time, temperature, storage conditions, duration, temperature, pressure)
3. The test method (protocol, extraction temperature, extraction duration)

A critical comparison of a number of methods described in literature has


been published elsewhere13 and we will only draw here the attention on a
116 Filled Polymers

Glass device for + Solvent Experimental


extraction kinetics at time ti method

Stopper
Extraction during
Solvent
a period ti+1 – ti
Steel wire cage
Collecting Vacuum
Sample at time ti+1 evaporation
of solvent
PTFE valve

Extract at
100 % BdR time ti+1 Weighing
initial gum content

residue
80
% Extract vs.

60 Extracted mass
at time ti+1
40 [%Extr]t = (100 – [%BdR])
20 ×[1 – exp(–bt)]
0
0 50 100 150 200 Bound rubber
Extraction duration, h at time ti+1

Figure 5.19
Extraction kinetic method for bound rubber assessment.

technique, based on the extraction kinetics, that offers the advantage to give
access to an “absolute” value of BdR (at the measurement temperature).22 As
illustrated in Figure 5.19, a special glass device is used to perform the swell-
ing of test samples in a fixed quantity of solvent for well-defined periods.
The sample, cut in small pieces, is weighed in a steel wire basket and dis-
posed in the glass vessel with a polytetrafluoroethylene valve at the bottom.
A know quantity of solvent (100 ml) is poured in the vessel and left in contact
with the sample; after various periods of time, the solvent (which contains
some extracted species) is collected through the valve and another portion
of pure solvent is introduced in the vessel for a further extraction period.
The procedure is repeated until complete extraction is achieved. Aliquot
quantities of collected extracts are evaporated under vacuum at 50°C and the
extracted quantity assessed by weighing the dry residue. Test data are thus
the extracted rubber (g) for various extraction periods (h).
Except for binary formulations (i.e., rubber  +  black), a correction is needed
for nonrubber soluble ingredients, generally the oil, the stearic acid, and
certain chemicals, by considering that at any time ti, those ingredients are
extracted in quantities proportional to the labile rubber. After this correction
is made, results are expressed in percent extracted rubber vs. the initial gum
rubber content in the compound, and used to draw the extraction kinetics
curve (see Figure 5.19). A model for the extraction kinetics is then used to fit
such data by non linear regression, according to:
Polymers and Carbon Black 117

[%extracted]t = (100 − [BdR]) × {1 − exp(−β t)} (5.7)

where [%extracted]t is the percent extracted rubber at extraction time t, [BdR]


the BdR content for an infinite extraction time, β an extraction kinetic param-
eter and t the time.
Figure 5.20 shows typical results obtained with this extration kinetic
method. CB filled compounds (N330, 50 phr) prepared with either high cis-
1,4 polybutadiene, ethylene-propylene rubber, or nitrile rubber, and stored
at room temparature (under cover) for 28 days, were extracted with toluene
at room temperature, according to the method described above. Extraction
data were fitted with Equation 5.7 and yielded the results given in Table 5.7.
As shown in Figure 5.20, quite long extraction periods are needed before an
extraction equilibrium is reached and how fast is this equilibrium reached
is clearly depending on the chemical nature of the rubber. Whatever is the
rubber however, Equation 5.7 has the capability to perfectly fit experimental
data and to yield significant and unambiguous BdR data.

100
BdR (%)
Cumultated extracted rubber, %

80 20.23
23.31
31.52
60
100
80
40 60
50 phr N330 cpds
40
20 EPDM
20 NBR
0 BR
0 10 20 30
0
0 100 200 300
Extraction time, h
Figure 5.20
Extraction kinetic data on N330 carbon black filled compounds; extractions in toluene at room
temperature; formulation: rubber, 100; carbon black, 50; ZnO, 5; stearic acid, 3; oil, 5; antidegra-
dants, 2; elastomers: EPDM, 57.5% ethylene, 4.7% ENB; NBR, 34.5% ACN; BR, 98% cis-1,4.

Table 5.7
Bound Rubber Content from Toluene Extraction Kinetic Method
Bound Rubber Extraction Kinetic
Rubber (% Initial Gum) Parameter β r2
EPDM 20.23 0.159 0.999
NBR 23.32 0.470 0.988
BR 31.52 0.176 0.994
118 Filled Polymers

As can be seen, the nature of the rubber significantly affects the BdR
c­ ontent, and also the extraction kinetics as expressed through the parameter
β. Such results do comply with contemporary theoretical views on the ori-
gin of BdR in CB filled systems, as a topological constraint effect, as will be
described hereafter. The extraction kinetic method does not only provide the
absolute BdR content (at the extraction temperature considered) with a de
facto compensation for experimental scatter, but also allows additional infor-
mation to be obtained by analyzing extracts, for instance by Gel Permeation
Chromatography (GPC). The residue can also be recovered, dried under
vacuum, and the BdR content crosschecked by thermogravimetry analysis
(TGA). The % BdR assessed through the extraction kinetic method is quite
reproducible (± 1%) when the test material is well mixed.
The above method is applicable only if the swollen rubber–filler gel remains
coherent during the extraction process, which is generally the case provid-
ing the filler level is high enough, i.e., higher than or close to the so-called
percolation level, ~12–13 vol.%, and the dispersion quality satisfactory, i.e., no
more agglomerates. When the swollen sample disintegrates, the extracted
solution is no longer clear and in extreme cases, it becomes very difficult, if
not impossible, to recover the rubber–filler complex.
In agreement with Figure 5.16, BdR can be viewed as a gel of (carbon) ­particles
with the bonding agent consisting of the longer polymer molecules. Such a con-
sideration is the root of many theories, intending to explain the very origin of
the phenomenon, with the goal to derive mathematical models that would allow
the BdR content of a given rubber formulation to be calculated from a set of fun-
damental material parameters. Most of those theories are consistent and lead
to useful conclusions (see a detailed discussion elsewhere13) but, in the author’s
opinion, ignore important aspects, such the chemical nature of the elastomer,
the (known) effects of mixing as well as storage effects.23–25 Figure 5.21 shows

Brabender EP2 mixer


40 Chamber volume = 90 cc OESBR 1712 - 50 phr N339
Rotor speed = 50 RPM
Temperature = 80°C
30
Bound rubber, %

20 OEBR 441 - 50 phr N339

10

0
0 5 10 15 20
Mixing time, min
Figure 5.21
Kinetic aspects in bound rubber formation. (    - Data from G.R. Cotten, Rubb. Chem. Technol,
57, 118 (1984); G.R. Cotten, Plast. Rubb. Proc. Appl., 8, 173 (1987).)
Polymers and Carbon Black 119

40
Compounds with 40 phr N330 and other usual
35 ingredients [Leblanc, Hardy (1991)]
Bound rubber (% initial gum) Natural rubber (SMR 5CV)
30

25

20 Polybutadiene (NeoCis BR40)

15

10 EDPM (Dutral TER054/E)

5 EPR (Dutral CO054)

0
40 60 0
80 20
100 120 140 160
Storage maruration* period (days)
* Storage at room temperature under dark plastic cover
Figure 5.22
Effects of elastomer and storage on bound rubber. (Data from J.L. Leblanc, P. Hardy, Kautsch.
Gummi, Kunstst., 44, 1119–1124, 1991; curves are drawn using a model described hereafter.)

for instance data from Cotten, which suggest that BdR does not form instan-
taneously during mixing, but according to a ­kinetics that will be discussed
later. Moreover, as illustrated in Figure 5.22 both the chemical nature of the
elastomer and the storage time (at room temperature) affect the BdR.
Let us consider that rubber–CB interactions have a pure physical nature.
Then such interactions are likely to be topological constraints exerted on
chain segments by the appropriate (geometrical) elements on the surface
of filler particles. This simplistic view was considered by Leblanc26 who
assumed that, from a strong interaction to occur, the surface topology of a
given CB particle must locally encounter the conformation of a chain seg-
ment equal to at least three structural units typical of the elastomer consid-
ered (Figure 5.23). This would be possible provided the polymer segment of
structural units and the filler topological site have the corresponding recip-
rocal geometry, in the appropriate orientation, and at the right time. During
mixing, the probability of such favorable events is obviously quite high. Once
this topological interaction has taken place, it is quite obvious that, in order
to release it, the free portions of the chain must exert on the constrained
units not only sufficient stresses but also in the appropriate direction; a pro-
cess that would require quite high energy level to be statistically significant
for BdR to vanish. The key point in this reasoning is that a logical estimate
can now be offered for the size of an “active” site on the filler particle, with
respect to the dimensions of the monomer unit; for instance it can be con-
sidered as two to three times the half lateral surface of a monomer unit,
easily calculated with C–C and C–H bonds length. The chemical nature of
the elastomer plays thus, its role through the length of the segment needed
for a strong interaction to occur, with chain unsaturation involved through
120 Filled Polymers

Rigid chain motif blocked by the Dangling segment


appropriate site on filter surface

C–C bond half-free Surface of carbon


rotation black particle

Dangling
segment Steps-like structure
at edges of broken
graphic plies

Graphitic layers

Figure 5.23
Rubber–carbon black interaction viewed as a topological constraints effect.

the associated local segment rigidity. An appropriate explanation for the


low BdR values obtained with ethylene-propylene rubber compounds is
in the meantime obtained, i.e., longer segments, ~C24, must conform their
morphology with the particle surface, an event which obviously has a lower
probability.
By somewhat extending an approach by Cohen–Addad for Silica-PDMS27
the variation of BdR vs. storage time (at constant temperature), can be con-
sidered as follows:

 M0 c Sp 
BdR(t) =  M n (0) + [ M n (∞) − M n (0)]( 1 − exp(−βt)) (5.8)
 A0 N Av 

where M0 is the weight of one skeletal bond, c the filler concentration (g/g of
gum polymer), Sp the specific surface area of CB (m²/g), A0 the ­average area
of one interactive site (nm²), NAv the Avogadro number, M n (0) and M n ( ∞ )
the number average molecular weight (g/mol) of chains involved in BdR,
respectively immediately at dump and after an infinite storage time, t the
time and β a kinetic parameter describing the storage ­maturation. Data in
Figure 5.22 were fitted with this equation.
The view that BdR is at best the adsorption–desorption balance of rub-
ber segments on CB sites at a given time (and temperature) is clearly
supported by the above model. As long as an equilibrium is not reached
however, the adsorption–desorption mechanism evolves while the com-
pound is at rest, hence the observed variations upon storage. How fast this
equilibrium BdR is reached depends on the chemical nature of the rubber,
on the compound formulation, on the mixing and the dump compound
storage conditions.
Polymers and Carbon Black 121

In the topological constraints effects behind rubber–filler interactions, there


is a competition between short and long chain segments. During ­mixing,
owing to the strong flow fields prevailing in the mixer, one considers that
all polymer fractions have equal probability to have segments arriving in
close contact with the appropriate sites on filler particles, in order to develop
a topological interaction. After dump, when the material is at rest, segments
from short chains are progressively displaced from active sites by segments
from long chains, leading to an increase in the average molecular weight of
BdR. This is exactly the mechanism (mathematically) expressed by the last
right member of Equation 5.8.

5.1.6  Viscosity and Carbon Black Level


How the addition of small solid particles to a liquid affects its viscosity (and
other properties) is a long lasting problem, with many theoretical attempts.
With respect to the above sections, that underlined specific aspects due to
CB nature and characteristics, it is clear that no global understanding can
yet be expected, that would be valid whatever the fluid and the filler. Let
us consider that the fluid is a polymer (melt); then the complexity of the
problem can nevertheless be stated in the form of a general functional, as
follows:

Pcpd = Ppolym × F ( Φ, d , S, X , T ,…) (5.9)

where Pcpd is the property (for instance the shear viscosity) of the mixture
polymer  +  filler, Ppolym the same property of the pure polymer, F() an appro-
priate mathematical function, Φ, d , S, X , T ,... the variables of this function,
respectively the filler volume fraction Φ, a parameter d related to particle
size, another one S describing the structure of the filler, then the rate of defor-
mation X and the temperature T, etc. All theoretical approaches to describe
the effect of filler fraction on compound viscosity can then be considered as
attempts to partially describe the functional F().
As a first approach, one may fix parameters S, X,  and T. In other terms, one
considers how the viscosity of a Newtonian fluid, in isothermal conditions,
is affected by the volume fraction of equal diameter spherical particles. With
ηpolym and ηcpd the viscosities respectively of the pure polymer and of the
filled compound, the following quantities are defined:

ηcpd − ηpolym
• The specific viscosity: ηsp =
ηpolym
ηcpd
• The relative viscosity (of the compound): ηrel =
ηpolym
It follows that ηsp = ηrel −1.
122 Filled Polymers

Historically, one of the earliest approaches was made by Einstein who con-
sidered a Newtonian medium [of viscosity ηpolym ≠ f ( X ) ] in which rigid
spherical particles of equal diameter are suspended. Providing the number
of particles is small enough for no interaction to occur between spheres, such
a suspension flows macroscopically in an apparently simple manner, when
submitted to a shear stress. However, when carefully observed on a finer
scale, it becomes obvious that, in the neighborhood of a particle, the flow
is not homogeneous because the liquid must flow around it. Thus, the local
shear rate of the liquid itself varies from point to point, and the average value
is larger than the overall rate of shear of the whole suspension. Consequently,
the global viscosity of the suspension is greater than the viscosity of the sus-
pending liquid. Einstein considered a suspension so dilute that no hydrody-
namic interaction occurs between different particles, and treated the liquid
as a continuum in which ordinary laws of hydrodynamics apply. He then
established the well-known equation:

ηcpd
ηcpd = ηpolym (1 + 2.5 Φ)    or   ηsp = − 1 = 2.5 Φ (5.10)
ηpolym

where Φ is the volume fraction of suspended (spherical) particles. With


respect to hypotheses in Einstein’s analysis, the applicability of Equation 5.10
is limited (very dilute suspension of rigid spherical particles) and ­several
authors,29–36 by using similar arguments, have considered other particle
geometries (ellipsoids, rods, disks) and/or slightly larger volume fraction, as
summarized in Table 5.8.
Owing to their mathematical simplicity, quadratic (and cubic) models (see
Figure 5.24) for rigid spheres suspension are attractive for low reinforcing
carbon blacks (e.g., N990) for which, for instance, the well known Guth, Gold,
and Simha equation fit well data up to a volume fraction equal to 0.20. The
cubic model by Vand was found to give good fit for glass spheres suspen-
sions (0.013 cm diameter) up to Φ = 0.37.
With reinforcing carbon blacks, the best spatial envelope for an aggregate
is the ellipsoid and, owing to its mathematical simplicity, models proposed
by Guth and Gold for spheres and rigid revolution ellipsoids drew the atten-
tion of several authors. For instance, White and Crowder37 considered the
following equation:

ηcpd = ηpolym (1 + 2.5 × f × Φ eff + 14.1 × f 2 × Φ eff


2 ) (5.11)

where f is an anisometry factor and Φeff an effective filler volume fraction that
takes into account the BdR, i.e., Φeff  = Φblack  +  ΦBdR, where ΦBdR is the ­volume
fraction of BdR.
Table 5.8
Effect of Fillers on (Newtonian) Viscosity—Theoretical Approaches
Type of Filler Low Volume Fraction Medium Volume Fraction
Rigid spheres with equal ηsp = 2.5Φ   (Einstein) ηsp = 2.5 Φ + 14.1 Φ 2
diameter (Guth, Gold, Simha)

ηsp = 2.5Φ + 12.6Φ 2 (Simha)

ηsp = 2.5 Φ + 7.17 Φ 2 + 16.2 Φ 3 (Vand)


Polymers and Carbon Black

Rigid revolution ellipsoids ηsp = 0.67 f Φ + 1.62 f 2 Φ 2    (Guth, Gold)

Rigid rods
 f   f   Kf 2  2
ηsp =  + 2  Φ    (Jeffery) ηsp =  + 2 Φ +  Φ    (Guth, Gold)
L  2 ln( 2 f)− 3   2 ln( 2 f ) − 3   ( 2 ln( 2 f ) − 3)2 
f=  : anisometry factor
D
K = constant, close to 1
L = large axis  f2 
ηsp =  2.5 +  Φ    (Huggins, Kuhn)
D = small axis  16 

 
 f2 
ηsp =   Φ    (Eisenschitz)
 15 ln(2 f ) − 45 
 2 

 
 f2 f2 14 
ηsp =  + +  Φ   ( Simha )
 15  ln(2 f ) − 3  5  ln(2 f ) − 1  15 
  2   2  

(Continued)
123
124

Table 5.8  (Continued)


Type of Filler Low Volume Fraction Medium Volume Fraction
Rigid disks
 4f 
ηsp =   Φ    (Jeffery)
 3 a tan( f ) 
L
f=  : anisometry factor 2
D  4f 
ηsp =   Φ    (Guth, Jeffery)
 3 a tan( f ) 

 16 f 
ηsp =   Φ    (Simha)
 15 a tan( f ) 
Filled Polymers
Polymers and Carbon Black 125

6
ηsp = 2.5 Φ + 14.1 Φ2 (Guth, Gold, Simha)
Specific (Newtonian) viscosity ηsp = 2.5 Φ + 12.6 Φ2 (Simha)
ηsp = 2.5 Φ + 7.17 Φ2+ 163 Φ3 (Vand)
4

2 ηcpd
ηsp = ηsp = 2.5 Φ
–1
ηpolym
(Einstein)
0
0 0.1 0.2 0.3 0.4 0.5
Volume fraction Φ
Figure 5.24
Effect of filler content on (Newtonian) viscosity; models for rigid spherical particles.

5.1.7 Effect of Carbon Black on Mechanical Properties


Once they have been shaped into a suitable object, rubber compounds are
vulcanized in such a manner that the full development of their mechanical
properties is achieved, without creep phenomena that are normally exhib-
ited by all polymer materials when on their rubbery plateau. In other terms,
vulcanization extends toward infinity the rubbery plateau of the relaxation
modulus function G(t). Furthermore, reinforcing fillers somewhat increase
the magnitude of the modulus at a given time. The combination of vulca-
nization and reinforcing effects induces quite complex changes in material
functions of polymers, as easily demonstrated through purposely simple
calculations.
Let us consider, for instance, the relaxation modulus function G(T) of a
pure SBR, as reported by Nielsen38 (Figure 5.25). The effect of CB loading (at
constant temperature) can be approached by rewriting the Guth, Gold and
Simha equation as follows:

Gcpd = Gpolym (1 + 2.5 × Φ + 14.1 × Φ 2 ) (5.12)

where Gpolym and Gcpd are the moduli, respectively of the pure polymer and
of the compound with a filler volume fraction Φ. The filler incorporation
effect can be simulated by calculating point by point Gcpd with several values
of Φ, for instance from 0 to 0.4. The Gcpd (T) functions obtained for all the
Φ considered are drawn vs. an arbitrary scale of the process. To simulate
changes imparted by vulcanization, we may consider that the modulus var-
ies from its actual value in the so-called flow (or terminal) zone, i.e., Gterm, to a
value corresponding to the rubbery plateau, i.e., Gvulc, using a simple sigmoid
function, for instance:
126 Filled Polymers

1e+4
1e+3
1e+2
Pa)

1e+1
Shear modulus (M

1e+0
1e–1
1e–2 Scale = 100
1e–3 Shear modulus
1e–4 function for
1e–5 vulcanized cpd
1e–6
1e–7 100
80 )
1e–8 ale
–60–40 60 y sc
r
–20 0 ra Modulus variation
Tem 20 40
r bit
40 a with vulcanization
per 60 80 20 e(
atu
re ( 100 s tag
°C) 0 s
120 es
r oc Modulus increase
P
due to filler addition
Scale = 0
Carbon black filled Shear modulus
SBR 1500 compound function for gum
rubber
Figure 5.25
Expected variation of modulus function during filler loading and vulcanization.

Gvulc − Gterm
G(t) = Gterm + α
(5.13)
 t
1+  
 τ

where G(t) is the value of the modulus for a given vulcanization time t, τ the
necessary (curing) time to obtain G(τ) = (Gterm + Gvulc)/2, and α a ­parameter
related to the vulcanization rate. The G(T) curves so obtained have been
drawn vs. the arbitrary scale of the process. As can be seen in Figure 5.25,
filler loading and vulcanization drastically modify the modulus function,
with the largest changes occurring during vulcanization.
As previously underlined (Chapter 3), CB reinforcement must be appre-
ciated with respect to a balance of properties, some of them ­modified in
antagonistic manner when filler loading is changed. This commands to
consider first how specific mechanical properties are modified through CB
addition.
Tensile modulus, and elongation are, by far, the easiest mechanical
properties to measure. If globally, ultimate tensile properties (i.e., stress
strength and elongation at break) are improved through reinforcing CB
addition, the magnitude of the effect depends on both the chemical nature
of the elastomer and the temperature. Figure 5.26 compares the effect of
Polymers and Carbon Black 127

30 30
Styrene-Butadiene rubber
NR gum
SBR gum
NR + 30 phr N330
SBR + 30 phr N330
Stress at break, MPa

Stress at break, MPa


20 20

10 10

Natural rubber

0 0
0 40 80 120 160 0 40 80 120 160
Temperature, °C Temperature, °C
Figure 5.26
Effect of temperature and carbon black on tensile stress at break.

30 phr N330 CB in NR and SBR vulcanizates.39 The tensile strength (TS)


of the unfilled NR compound varies with temperature, in such a man-
ner that a drastic decrease is observed at around 80°C. This behavior is
explained by a strain crystallization effect of NR below this temperature.
It follows that the effect of a reinforcing CB depends on temperature.
The addition of 30 phr N330 increases the TS by 20% when the tempera-
ture is lower than 80°C, and by 200% above 80°C. SBR does not exhibit
strain crystallization effects and therefore only a slight decrease of the TS
with temperature is observed. A room temperature, adding 30 phr N330
increases the TS by a factor of 10, and only doubles it when the tempera-
ture is above 120°C.
Tensile stress softening (TSS), the so-called Mullins effect,40 is an impor-
tant phenomenon with CB filled rubber compounds, which describes the
fact that all elastomers, either vulcanized or not, either filled or not, require
a higher stress at equal strain during the first extension than during fur-
ther ones. As schematically described in Figure 5.27, a vulcanized rubber
sample is first stretched until an elongation ε1 is reached, then the stress is
released. From ε1 to ε = 0, the stress–strain curve is below the trace recorded
during the first extension, and there is a residual strain. The first extension
produces much larger changes in the stress–strain curve than subsequent
extensions, in such a manner that after three to four cycles, traces are practi-
cally superimposed.
A common explanation for the Mullins effect is that, between the first and
the second extension, the structure of the material has changed and it is gen-
erally observed that the higher the CB loading, the larger this effect. TSS is
not solely due to the presence of the filler, whilst the latter enhances the effect.
When a vulcanized rubber sample, which has been stretched twice, is sub-
mitted to a thermal treatment (for instance several hours at 80°C), it recovers
128 Filled Polymers

20
Extension : 1 2
15
Stress, MPa NR gum

10

0
0 1 2 3 4 5 6 7 8
Strain
20
Extension : 1 2 “Mullins” effect
15 2nd and further
Stress

NR/N330 extensions
Stress, MPa

(60 phr)
1st extension
10

ε1
5 2nd extension
Residual Elongation
0 strain
0 1 2 3 4
Strain
Figure 5.27
Tensile stress softening (Mullins effect).

essentially the first strain behavior, at least with certain types of vulcaniza-
tion systems (mono-sulfide systems or peroxide curing). This means that no
rupture process is involved in the TSS effect, which is therefore attributed to
a quasi-irreversible rearrangement of the molecular network, because local
deformations are not affine.
By extending the works of Guth, Gold, and Simha on viscosity–filler frac-
tion effects, one can express the effect of CB loading on modulus with a simi-
lar expression, but using an appropriate anisometry factor to somewhat take
into account the nonspherical shape of CB aggregates, i.e.:

Ecpd = Epolym ( 1 + 2.5 × f × Φ + 14.1 × f 2 × Φ 2 ) (5.14)

This approach implicitly considers that CB effect on modulus is essentially


hydrodynamic. The anisometry factor (or form factor) reflects the ratio large
axis/small axis of a revolution ellipsoid, considered as the best envelope for
an aggregate. When using f = 6, it is found that Equation 5.14 fit well experi-
mental data with N330 CB up to Φ = 0.3. Note that alternative equations have
been proposed, for instance by Guth:

Ecpd = Epolym (1 + 0.67 × f × Φ + 1.62 × f 2 × Φ 2 ) (5.15)


Polymers and Carbon Black 129

If however, one considers f as an adjustable parameter and one uses


equation Equation 5.14 (or Equation 5.15) to fit (using nonlinear regres-
sion algorithm) modulus data derived from first and second stress strain
curves of a Mullins type experiment, then one notes that the parameter
f has changed between the two stretching. This suggests considering
that there is a kind of transient rubber–CB structure which is affected by
the strain.

Medalia proposed another approach, through the equation:

Ecpd = Epolym (1 + 2.5 × Φ eff + 14.1 × Φ eff


2 ) (5.16)

where Φeff is the effective filler volume fraction, considered either with
respect to BdR content,41 i.e.,

 1 + 0.02139 × DPBA 
Φ eff = Φ black   (5.16a)
 1.46
or using:

Φeff = Φblack + ΦBdR (5.16b)

where DBPA is the dibutylphthalate adsorption number of the CB con-


sidered. Medalia proposed Equation 5.16a on the basis of his concept of
equivalent spherical envelope for CB aggregates. The constants in this
equation have been derived from Transmission electron microscopy
(TEM) observations on a selection of ­carbon blacks, combined with sev-
eral reasonable hypotheses. At best the Φeff so calculated is an estima-
tion and does not take into consideration the mere fact that BdR level is
strongly depending on the rubber nature. It is therefore always preferable
to measure BdR and to use Equation 5.16b. A similar but slightly differ-
ent equation, i.e., Φeff = Φblack[(1 + 0.0181 × cDPBA)/1.59] was proposed when
“crushed DBPA” is used to characterize the CB.43
In order to explain the effect of CB on mechanical properties, Mullins intro-
duced an interesting concept: strain amplification (Figure 5.28). At micro-
scopic level, a filled rubber compound consists of “hard regions”, i.e., where
CB concentration is locally high, and of “soft regions”, i.e., with locally low
black level. Such a view, in fact readily seen by microscopy with a suitable
magnification, corresponds very well with the fact that a part of the elasto-
mer in unvulcanized mixes is extractable with a good solvent, and therefore,
is not interacting with the filler. It follows that, when a stress is applied to
a vulcanized compound, soft regions essentially support the overall strain,
before hard ones are affected (in which case the TSS effect is observed). The
presence of CB particles brings therefore, an amplification of the average
strain supported by the elastomer.
130 Filled Polymers

Unfilled vulcanizate Filled vulcanizate Strain amplification factor


Soft phase
×2 Stretching (elastomer) ×2 Stretching
Effective strain
2.000 2.429 λ´
=X
2.000 Hard phase 1.000 λ
10 10
" (carbon black) Measured strain
" 7.285

Non reinforcing carbon black (spherical)


1.000
X=1 + 2.5 Φ + 14.1 Φ2
4.857

Only the soft phases 1.000 Reinforcing carbon black (anisometrical)


support the strain 2.429 X=1 + 0.67 f Φ + 1.62 f 2Φ 2

The strain supported by the elastomer phase (empirical)


is larger in the filled than in the unfilled material an isometry parameter
Figure 5.28
Concept of strain amplification in filled elastomers.

By considering an ideal network, the elasticity theory provides a relation-


ship for stress dependence upon strain, found applicable at (very) low strain,
i.e.:
E0  1  1
σ= λ − 2  = NkT  λ − 2  (5.17)
3  λ   λ 

where E0 is the Young’s modulus, N the fraction of active elastic chains in


the network, k and T the Bolztmann constant and the temperature respec-
tively and λ the strain (λ = L/L0; L0 = initial length of test sample; L = stretched
length). For larger strain (i.e., up to λ ≈ 3.0), the semiempirical equation by
Mooney and Rivlin is applicable, i.e.:

 C   1
σ = 2  C1 + 2   λ − 2  (5.18)
 λ  λ 

where the constant C1 and C2 (to be determined experimentally) are


­respectively a function of the fraction of active elastic chains and of the
type and concentration of CB, as demonstrated by experimental data
(Table 5.9).44
As can be seen, C1 does not depend on the type and level of CB, con-
trary to C2. In fact, the C2 term in the Mooney–Rivlin equation somewhat
expresses the deviation with respect to the ideal elastic network (as readily
seen when comparing Equations 5.17 and 5.18). When using this equation
Polymers and Carbon Black 131

Table 5.9
Effect of Carbon Black Type and Level on C1 and C2 Constants in Mooney–Rivlin
Equation
Filled Carbon Black Compounds Effect of Volumic Fraction ϕ
SBR 1500 100.00 Black ϕ C1 (MPa) C2 (MPa) λmax
Carbon black variable – 0 0.13 0.21 3.0
ZnO 5.00 N330 0.05 0.15 0.28 2.5
Stearic acid 1.00 0.10 0.16 0.45 2.0
Sulphur 1.75 0.15 0.17 0.62 1.8
CBS 1.00 0.20 0.20 0.86 1.6
Effect of Black Type at ϕ = 0.20 N550 0.05 0.15 0.24 2.5
Black C1 (MPa) C2 (MPa) λmax 0.10 0.15 0.43 1.8
N220 0.20 0.96 1.5 0.15 0.16 0.63 1.6
N330 0.20 0.86 1.6 0.20 0.20 0.74 1.7
N550 0.20 0.74 1.7 N762 0.05 0.15 0.23 3.0
N660 0.21 0.67 1.8 0.10 0.16 0.60 2.0
N770 0.26 0.57 1.8 0.15 0.16 0.52 1.8
0.20 0.18 0.61 1.6

in the case of a system filled with CB, it is implicitly accepted that the mean
effect of the filler consists in increasing the effective strain of the elastomer
phase. And because CB aggregates are rigid, the local strain of the rubber
matrix is larger that the overall measured deformation.
Within identical validity limits, Mullins and Tobin45 have shown that the
stress–strain behavior of black-loaded rubber vulcanizates corresponds to
the stress–strain response of pure gum vulcanizates multiplied by a suit-
able strain amplification factor X, which expresses the fact that the aver-
age strain supported by the rubber phase, is increased by the presence of
filler. In other terms, the effective strain of the elastomer matrix λ′ is given
by: λ′ = λ × X, where λ is the overall measured deformation of the filled
material.
The strain amplification factor is obviously depending on the filler loading
and likely on the filler characteristics, firstly the structure. For a nonreinforc-
ing CB (i.e., N990), which consists essentially of spherical ­particles with mean
diameter of about 400 nm, Mullins and Tobin showed that an appropriate
equation for the strain amplification factor is:

X = 1 + 2.5 × Φ + 14.1 × Φ 2 (5.19)

However, for a finer reinforcing black, a more suitable equation is:

X = 1 + 0.67 × f × Φ + 1.62 × f 2 × Φ 2 (5.20)


132 Filled Polymers

where f is an empirically determined value for the CB grade considered.


For instance f ≈ 0.65 was found suitable for N330, which consists of aggre-
gates of up to 300 nm (longest dimension). Both equations are essentially
empirical, with an obvious implicit reference to the works of Guth, Gold,
and Simha.
Recently, advanced Atomic Force Microscopy (AFM) techniques have
provided quite a ­convincing demonstration that in unfilled NR vulcani-
zates, tensile deformation is not homogeneous;46 in other words, the defor-
mation is far to be affine. It is fairly obvious that if nonaffine deformation
is observed on unfilled systems, ­nonhomogeneous deformation is surely
the rule with filled rubber vulcanizates. Of course the “strain amplifica-
tion” concept of Mullins and Tobin somewhat recognizes the macroscale
occurrence of such inhomogeneities and, moreover, the BdR concept is
also an explicit demonstration that, close to CB particles, there are rubber
segments that are immobilized, so that they are in a pseudo-glassy state.
Except maybe at excessive strain, the tightly BdR is not expected to support
the strain, so that only the loosely BdR and the extractible rubber ­support
it. But of course the stress is mainly supported by the filler particles and the
tightly BdR. Local nonhomogeneous deformation is thus the rule in rubber
materials, so that all theoretical considerations that either implicitly or by
hypothesis, consider affine deformation are bound to fail or, at least, to have
a very narrow applicability range. The TSS softening effect, when observed
with an unfilled rubber vulcanizate, is thus attributed to the nonaffine
displacement of the junction points of the network and to the incomplete
recovery of their original positions after the strain is back to zero. It follows
that the TSS effect is larger in CB filled systems (see Figure 5.27 for instance)
because of the strain amplification effect.
It must be noted here that other explanations have been proposed for the
reinforcing effect of CB. Blanchard47 considered that CB strongly modifies
the nature of the elastomer network with, amongst other features, strong
and weak links between rubber segments and appropriate sites at the sur-
face of particles. Recent considerations are close to this concept, as we shall
see later.
Up to this point, the strain amplification factor can be viewed as a mere
empirical approach to assign the modulus increase in CB filled compound
to filler level. Equation 5.19 above essentially resulted from considerations
on the hydrodynamic effects induced by the presence of solid particles
ideally dispersed in a matrix with a considerably lower modulus. The
empirical factor f in Equation 5.20 adds nothing in this respect and it is
well known that both equations do not suit at all either highly loaded
compounds, whatever is the grade of CB, or moderately loaded materials
with high structure blacks. Over the last decades, several authors have
developed theoretical considerations to model the likely effect of a so-
called filler network structure and the associated energy dissipation pro-
cess when filled compounds are submitted to increasing strain.
Polymers and Carbon Black 133

Earlier works by Medalia and coworkers demonstrated that certain key


mechanical properties of carbon filled compounds were related to the prod-
uct of the filler volume fraction × the overall rubber-filler interfacial area per
unit volume of compound.48,49 More recently, Wang et al.43 observed a good
correlation between dynamic properties, namely tan δ, and an ­interaggregate
distance, which was calculated with respect to a random distribution of
equivalent spheres and the concept of occluded rubber. They concluded
therefore that filler particle attractive forces and interaggregate distance
were controlling factors for the so-called filler network and the resulting
enhancement of compound’s modulus.
A CB filled rubber compound can therefore be considered as a kind
of soft three-dimensional network in which aggregates would act as
“anchoring knots” for several elastomer chains. One part of the rubber
would consequently plays a particular role in connecting CB aggregates,
in other words, in making “junctions” between them. Such consider-
ations are obviously well in agreement with the concept of “BdR,” as dis-
cussed above (Section 5.1.3). By paying attention to this “junction rubber,”
Ouyang et al.50,51 developed an interesting approach, generally referred
as the network junction (NJ) theory, in which the interaggregate distance
is the central argument. There are unfortunately a number of misprints
and unit inconsistencies in the published equations, which prompted the
author to reconsider the theoretical development in parallel with para-
metrical verifications (see details in Appendix 5.1).
Let us consider a given mass MCB of a carbon black at the end point of
the so-called “crushed” dibutyl phtalate adsorption test (ASTM D-3493;
see Section 4.1.3 above). At this stage, all the aggregates of the CB sample
are likely exhibiting the most compact arrangement in a DBP matrix and,
with respect to the overall volume of the DBP + CB mixture, the filler vol-
ume fraction is maximum and can be assessed from:

MCB
ρCB
Φ maxCDBP = (5.21)
MCB
+ cDBP ⋅ MCB
ρCB

where ρCB is the specific gravity of the CB and cDBP its crushed DBP adsorp-
tion number (cm3 DBP/100 g of filler). As we have seen (Section 4.1.4, Equation
4.7), the solid volume of an aggregate is given by: Vs = Np(πd3/6), where Np is
the number of elementary particles of diameter d. The number of aggregates
NaCB in MCB grams of CB is thus:

MCB
NaCB = (5.22)
π d3
Np ρCB
6
134 Filled Polymers

and therefore, the number of junctions (or contact) points between


­aggregates is NaCB(ς/2) if ς is the average number of contact points between
­neighboring aggregates. If one makes the (strong) hypothesis that in a rub-
ber compounds, dispersed aggregates keep the same compact arrange-
ment as in DBP, except that the distance between neighboring aggregates
is expanded by a distance hgap, and if a cubic arrangement of aggregates is
considered for the sake of simplicity (Figure 5.29), it comes that the CB vol-
ume fraction is given by:

MCB
ρ
Φ CB = 3
(5.23)
  MCB 
1/3
 ζ 
1/3
+ CDBP ⋅ M + h ⋅
gap  Na ⋅
  ρ

CB 
  CB 2  

Compact cubic arrangement d


Compact cubic arrangement aggregates in rubber matrix
of carbon black aggregates
(cDBP absorption test)
D

hgap

Aggregates of Np
elementary particles

150 150
100 Particles
200 Particles
Junction gap width, mm

Junction gap width, mm

300 Particles
100
N330, Ф = 0.20
100
50
2 Junctions
5 Junctions
10 Junctions
0 0
0 5 10 0 100 200 300
Number of contact points Number of primary particles/aggregate

Figure 5.29
Network junction model; the two graphs were calculated with Equation 5.25 and the following
data: CTAB = 41.32 m2/g, ρCB = 1.86 g/cm3, Φmax = 0.38, ΦCB = 0.20.
Polymers and Carbon Black 135

In the rubber compound, the maximum CB volume fraction is thus, lower


than in the mixture with DBP because rubber chains move apart neighbor-
ing aggregates by the distance hgap. An expression can then be derived for the
junction distance, i.e.:

− ( 1/3 )
 3ς   1 1 
hgap = d3 − (5.24)
Φ max 
3
 π N p   Φ CB

where ς is the average number of contact points between neighboring


aggregates, Np the number of elementary particles of diameter d in an
aggregate, ΦCB and Φmax, respectively the filler volume fraction and the
maximum packing fraction of the CB grade in the rubber compound. How
the junction gap width varies with both the number of contact points and
the number of primary particles per aggregate is illustrated by the graphs
drawn at the bottom of Figure 5.29. If one considers that an aggregate con-
sists of a simple assembly of touching elementary particles, one has the
following relationship between the specific surface area (in m 2/g) and the
elementary particles diameter d: Ssp = 6/dρ. Equation 5.24 can consequently
be rewritten as:

− ( 1/3 )
 3ς  6  1 1 
hgap =  − (5.25)
ρ Ssp  Φ CB Φ max 
3 3
 π N p 

So, according to this approach, in a given compound, the junction gap


hgap is inversely proportional to the specific surface area of the CB, as mea-
sured for instance by the adsorption of CTAB (ASTM D3765; see Section
4.1.3 above). In addition, higher structure (i.e., higher CDBP adsorption)
blacks will give shorter junction gap because the Φmax would be smaller.
To readily use Equation 5.25, one needs characterization data on CB,
namely the specific area, the (average) number of particle per aggregate
and the number of junctions per aggregate. Characterization data as
compiled from various sources were previously given in Table 5.5 and
Np can be estimated with Equation 4.10 (Section 4.1.4). Providing a fair
hypothesis is made about ς, for instance six junctions per aggregate with
an explicit reference to a simple cubic arrangement of aggregates, then
the data in Table 5.10 can be calculated (see details in Appendix 5.1). Note
that the maximum volume fraction as calculated from CDBP adsorption
(Equation 5.21) was used.
In agreement with the theoretical considerations above, the more rein-
forcing the CB the smaller the junction gap width. It is worth noting that a
diene rubber with a molecular weight in the 200–400,000 g/mole range has a
136 Filled Polymers

Table 5.10
Junction Gap Width as Calculated with Equation 5.25 for 20% Volume Fraction
Carbon Black and Six Junctions per Aggregate
Elementary Particle
Diameter Number of Particles per hgap
Carbon Black (nm) Aggregate Φmax (nm)
N110 18.32 278 0.354 28
N220 20.95 285 0.354 32
N330 30.50 209 0.382 44
N550 50.58 298 0.390 100
N660 59.28 170 0.424 101
N774 87.20 97 0.460 114
N990 294.90 26 0.592 286

random coil diameter in the 20–40 nm range. It means that for the reinforcing
CB grades, one macromolecule is sufficient to establish a junction between
two neighboring aggregates. For less or no reinforcing grades, the junction
will always involve entangled rubber chains with, as a consequence, lower
reinforcing effects.
A relationship for the strain amplification factor can be derived from
the NJ theory by considering the analysis made by Gent et al.52,53 of the
behavior of a rubber volume bonded between two rigid spheres. Indeed
when a compressive (or a tensile) force F provokes a small displacement ∆x
of one sphere with respect to the other, there is a compression (or tensile)
­stiffness which consists of two parts: (1) the rubber layer compressed (or
stretched) between the two spheres, (2) the restraints at the bonded sur-
faces of the spheres. Gent and Park developed the following theoretical
equation, i.e.:

F π  h  D + h  2 D2 + D h − 2 h 2 
= ⋅  1 +  ⋅ ln  +  (5.26)
E0 ⋅ D ⋅ ∆x 8  D  h  2 h (D + h) 

where E0 is the modulus of the rubber, D the diameter of the spheres and h
the initial gap. Experimental data, as well as finite-element method results
confirm the validity of the model but its mathematical form makes it not easy
to handle. However, Equation 5.26 essentially predicts an inverse depen-
dence of the stiffness upon the ratio h/D for relatively large separation, i.e., in
a layer thickness range down to one-tenth of the sphere radius. When using
logarithmic scales, such a dependence appears to correspond to a straight line
with a negative slope. It is therefore, attractive to approximate Equation 5.26
with a simple power law (see Appendix 5.1 for details). Using data ­published
Polymers and Carbon Black 137

by Gent et al., it can indeed be shown that a much simpler equation can be
considered, i.e.:
−β
F π  h
= α ⋅ ⋅  (5.27)
E0 ⋅ D ⋅ ∆x 8  D

with α ≈ 1.22 and β ≈ 0.96.


If one considers a volume V of filled compound, the overall number of CB
aggregates is N/V = ΦCB × ρCB × NaCB/MCCB, where MCB is the mass of CB in
the compound that corresponds to the volume fraction ΦCB. With respect
to Equation 5.22, it follows that in a cross sectional area of the sample (i.e.,
2
 3 V  ), the number of aggregates is:

2/3 2/3
 N  6 Φ CB  1
 V  = ⋅ (5.28)
 π N  p d2

When the sample is stretched (of compressed) along one axis by a strain
∆X, there will be (N/V)2/3 junction gaps contributing to the stress, in addi-
tion to the response of the unfilled rubber fraction. With respect to Equation
5.27, the junction gaps contribute to a force over the cross section V 2/3 and,
because obviously ∆X = ∆x ⋅ (N/V)1/3, the NJ contributes to the modulus by a
quantity:
1/3 −β
FJ  N π  hgap 
= ⋅ E0 ⋅ Des ⋅ α ⋅  (5.29)
∆X  V  8  Des 

where Des is the equivalent sphere of an aggregate, with reference to Medalia’s


floc simulation (see Section 4.1.4). When expressing Des in terms of number of
elementary particles and substituting Equations 5.24 and 5.28 into Equation
5.29, it comes:
( 1+β )/3 −β
FJ π  6 Φ CB   ζ
β/3
  Φ CB  1/3 
ε (β + 1)/2
= E0 ⋅ α ⋅ ⋅  ⋅N 2
⋅  ⋅ 1 −    (5.30)
8  π N p 
p
∆X  2   Φ max  

If one considers that, over the cross section, there is also a contribution
from the rubber matrix alone, then the associated stiffness is estimated as:

FR

∆X
= E0 ⋅ ( 1 − Φ 2CB
/3
) (5.31)

The strain amplification factor Xf is eventually given by:

( 1+β )/3 −β
π  6 Φ CB   ζ
β/3
  Φ CB  1/3 
B ) (5.32)
+ ( 1 + Φ 2CB
ε (β + 1)/2
Xf = α ⋅ ⋅  ⋅N ⋅  ⋅ 1 −   
/3
8  π N p 
p
 2   Φ max  
138 Filled Polymers

Equation 5.32 is quite different from the equation published by Ouyang for
at least two reasons: first we have used the Medalia’s aggregate equivalent
sphere diameter in exploiting the two spheres model of Gent et al., second
misprints may be suspected in Ouyang’s publications because several of his
equations are suffering from unit inconsistencies.
It is worth comparing the effectiveness of Equation 5.32 both in meet-
ing experimental data and with respect to the well known Guth and Gold
approach, as considered by Mullins and Tobin. In the author’s opinion,
such a comparison is all the more valid if data from the literature are
used. Unfortunately not many full sets of data are available with all the
necessary information to perform such an exercise. The graphs in Figure
5.30 were drawn using G′ data from Payne and Whitakker39 on butyl rub-
ber compounds with HAF CB (that we assimilated to N330, in order to
calculate Np and Φmax with the average DBP and CDBP data from Table
5.5), E′ data from Caruthers et al.48 on SBR1500 compounds with N347
(from which we extrapolated the unfilled compound value, not given by
the authors) and G′ data from Wang.54 In using Equation 5.32, the param-
eters α and β were respectively taken as 1.223 and 0.964 and two junctions
per aggregate were considered (i.e., ζ = 2) ; details of the calculation are
given in Appendix 5.1.
As can be seen, when compared with the Guth and Gold equation, the
NJ model predicts generally a sharper strain amplification factor as CB
volume fraction increases, but fits better the ­experimental data in the case
of SSBR compounds data. In applying Equation 5.32 however, parameters
α, β, and ζ were fixed and the only CB dependent parameters were Np
and Φmax, readily calculated from available DBP and cDBP data, respec-
tively. As shown in Figure 5.30b, when using the maximum theoretical
volume fraction for a cubic arrangement of spheres of equal ­d iameter (i.e.,
Φmax = 0.524), and only playing with the value of Np considerably improves
the manner the model is meeting experimental data. The model is not
perfect but, in the author’s opinion, the above comparison nevertheless
demonstrates the validity of the NJ theory, particularly with respect to all
the assumptions and simplifications made. Indeed, in a given CB sample,
aggregates are varying in shape and size and do not conform really to
equal diameter spheres. The effective maximum packing fraction for a
given black is thus, likely different. Similar comments apply to the cal-
culated number of particles per aggregate but in a lower extent because
when considering either 100 or 300 particles/aggregate the terms involv-
ing Np in Equation 5.32 vary from 0.50 to 0.42. Finally in the calculations
above, considering only two junctions per aggregate was an arbitrary
choice; using six instead of two would give around 50% more weight to
the associated term.
Polymers and Carbon Black 139

(a) 10
Butyl rubber/N330 cpds
G0́ = 0.2 MPa
G´ modulus, MPa

Guth & Gold equation


Network junction model
5 Np = 209
Φmax = 0.390

0
0 0.1 0.2 0.3
Carbon black volume fraction

(b) 40
SBR1500/N347 cpds
E0́ = 4.625 MPa
E´ modulus, MPa

Guth & Gold equation


Network junction model
20 Np = 333
Φmax = 0.357
Network junction model
Np = 200
Φmax = 0.524
0
0 0.1 0.2 0.3
Carbon black volume fraction

(c) 20
SSBR/N234 cpds
G0́ = 0.769 MPa
G´ modulus, MPa

Guth & Gold equation


Network junction model
10 Np = 348
Φmax = 0.347

0
0 0.1 0.2 0.3
Carbon black volume fraction
Figure 5.30
Comparing the strain amplification factor as calculated from the network junction theory with
experimental data and the Guth and Gold equation: Butyl rubber compounds data are from
A.R. Payne, R.E. Whittaker, Rubb. Chem. Technol., 44, 440–478, 1971; SBR1500 compound data are
from J.M. Caruthers, R.E. Cohen, A.I. Medalia, Rubb. Chem. Technol., 49, 1076–1094, 1976; SSBR
compounds data are from M.J. Wang, Rubb. Chem. Technol., 71, 520–589, 1998; For the three car-
bon blacks considered, average DBP and cDBP data from Table 4.5 were used in calculating the
number of elementary particles per aggregate and the maximum filler fraction.
140 Filled Polymers

5.1.8 Effect of Carbon Black on Dynamic Properties


The significant effects of CB on the dynamic properties of elastomers are
­particularly important with respect to tire technology. Indeed carbon blacks
(as well as other type of fillers) modify dynamic modulus functions in
regions which are particularly relevant for the tire behavior and therefore
the driving comfort and safety of an automotive vehicle. Let us consider the
typical dynamic functions G′(ω) and G′′(ω) of a rubber material (Figure 5.31).
The rolling resistance of a tire is depending of the relative magnitudes of the
elastic and viscous moduli in the 100–1000 Hz frequency range; the lower
the tan δ, the lower the viscous dissipation and hence the rolling resistance.
But the wet skid resistance of a tire tread is depending of the relative G′ and
G″ in the 100–10,000 KHz region in such a manner that the higher the tan δ,
the safer the tire. In other terms, improving both the rolling and the wet skid
resistance requires to play with the dynamic moduli in antagonistic manner
with respect to the frequency range concerned. Many attempts were made
in the 1970–1980s by modifying the (structure) of the elastomer, through
advanced synthesis techniques, without real success however as it proved
not really possible to induce the desired antagonistic changes in the tan δ vs.
frequency function by only changing the macromolecular architecture. The
solution came from a (radical) change in the nature of the filler, i.e., from CB
to (silinated) silica, as we will see later.
How carbon blacks do affect the dynamic properties in both the rubbery
plateau and the transition zone is therefore a problem with tremendous tech-
nological implications, particularly when “large” amplitude strain are con-
cerned (in practice when γ  >  0.01 [i.e., 1%]).

Terminal Rubbery Transition Glassy


10 zone plateau zone zone
Log G´´(Pa)

0
GN

6
Log G´ (Pa)

G΄΄
4
Log ω (Hz)
–5 0 5 10
Region related with
rolling resistance Region related with
of a tire wet skid resistance
of a tire tread
Figure 5.31
Dynamic properties of rubber materials with respect to tire technology requirements.
Polymers and Carbon Black 141

As may be expected, dynamic properties are affected in a complex ­manner


by CB type and level, frequency and temperature. No overall understanding
is presently available, but a certain coherence emerges when reviewing the
most significant effects of carbon blacks on dynamic properties, as reported
in literature.

5.1.8.1 Variation of Dynamic Moduli with Strain Amplitude


(at Constant Frequency and Temperature)
The general observation, reported by many authors, is that the addition of
CB to an elastomer induces typical non-linearity in the viscoelastic behavior.
Figure 5.32, drawn using data on SSBR compounds published by Wang,54
shows typical variations of both G′ and G′′ with strain amplitude (at constant
frequency and temperature). Qualitatively the same observation is made
on many filled systems, whatever the chemical nature of the rubber or its
macromolecular characteristics, and irrespective of the grade of CB used. At
very low strain, the elastic modulus G′ increases with increasing CB fraction
and there is little or no sensitivity on strain amplitude, a characteristic of
the linear viscoelastic behavior. As strain increases a drop in G′ is observed
and the higher the filler content the larger this drop. This effect is typical of
many filled systems and is referred as “dynamic stress softening (DSS)” (or
“Payne” effect, with regard to the important contribution of this author). It
is also observed that the viscous modulus G′′ passes through a maximum

N234 content
Φ phr SSBR Duradene 715 100
12.0 0.000 0 N234 carbon black Variable
Elastic modulus G´, MPa

Zinc oxide 3
0.124 30 Stearic acid 2
9.0 0.159 40 Antioxidant 1
Sulphur 1.75
6.0 0.191 50 CBS 1.25
0.220 60 MBT 0.2
0.248 70
3.0
1.6
Viscous modulus G´´, MPa

0.0
10–1 100 101 102 1.2
Double strain amplitude, %
0.8
Dynamic spectrometer II
(Rheometrics Inc.)
Temperature : 70°C 0.4
Frequency : 10 Hz
0.0
10–1 100 101 102
Double strain amplitude, %
Figure 5.32
Strain sweep experiments on SSBR compounds with various levels of carbon black.
142 Filled Polymers

value as the strain amplitude increases and in line with the G′ behavior, the
larger the CB content, the higher the G′′ peak.

5.1.8.2 Variation of tan δ with Strain Amplitude and


Temperature (at Constant Frequency)
At constant filler level, the magnitude of DSS is depending on the CB struc-
ture as shown in Figure 5.33. Either at 0 or at 70°C, an unfilled (gum) solu-
tion SBR vulcanizate shows no sensitivity to strain amplitude up to 40%
strain; the only effect of a temperature increase is to lower both G′ and G′′
and hence tan δ. With CB, the dynamic behavior is more complicated and
strongly depends on the reinforcing character of the filler; at low tempera-
ture (i.e., 0°C), the more reinforcing the CB, the lower tan δ at low strain.
As strain amplitude increases the black structure effect is reduced, until the
large (nonlinear) drop is observed. At higher temperature (i.e., 70°C), the
reverse effect is observed: the more reinforcing the CB the higher the tan δ
peak occurring at around 10% double strain amplitude.

5.1.8.3 Variation of Dynamic Moduli with Temperature


(at Constant Frequency and Strain Amplitude)
Generally the elastic modulus of filled rubber vulcanizates decreases with
increasing temperature, but the magnitude of the variation is strongly
depending on the strain amplitude, as illustrated in Figure 5.34, in the case of

0.70 SSBR Duradene 715 100


N234
10 Hz; 0°C Carbon black 50
N347 Zinc oxide 3
0.60 N660 Stearic acid 2
Gum Antioxidant 1
Sulphur 1.75
Tangent δ

0.50 CBS 1.25


MBT 0.2
0.40 0.24
10 Hz; 70°C
0.20
0.30
0.16
Tangent δ

0.00
0 1 10 100
Double strain amplitude,% 0.12

0.08

Dynamic spectrometer II 0.04


(Rheometrics inc.)
0.00
0 1 10 100
Double strain amplitude,%
Figure 5.33
Effect of carbon black type and temperature on tan δ of SSBR compounds.
Polymers and Carbon Black 143

15
20°C
Natural rubber
Elastic modulus G´, MPa N330 (32% vol.)
40°C
10
60°C

5 90°C

0
0.001 0.01 0.1 1.0
Double strain amplitude
Figure 5.34
Effect of temperature on the elastic modulus.

104 N234 content SSBR Duradene 715 100


Φ phr N234 carbon black Variable
Elastic modulus G´, MPa

103 Zinc oxide 3


0.000 0 Stearic acid 2
0.045 10 Antioxidant 1
102 0.124 30 Sulphur 1.75
0.190 50 CBS 1.25
101 0.248 70 MBT 0.2
2.0
100
1.0
10–1 0.5
–60 –40 –20 0 20 40 60 80 100 120
Tangent δ

Temperature, °C 0.3

Dynamic spectrometer II 0.1


(Rheometrics Inc.)
Double strain amplitude : 5% 0.05
Frequency : 10 Hz 0.03
–60 –40 –20 0 20 40 60 80 100 120
Temperature, °C
Figure 5.35
Effect of carbon black level and temperature on dynamic properties.

a NR compound with 50 phr N330 CB. As can be seen, at (very) low strain, the
elastic modulus is roughly divided by two when the temperature increases
from 20 to 90°C. For larger strain amplitudes, quite common in fact in many
practical applications, the temperature effect becomes negligible.
Changing the CB content further complicates the temperature effect. As
shown in Figure 5.35 in the case of solution SBR compounds with vari-
ous levels of CB, G′ slightly increases with CB level at low temperature.
Above a certain temperature, a severe drop is observed but increasing
CB level somewhat dampens the temperature effect. For instance from
144 Filled Polymers

−40 to + 40°C, the vulcanized gum compound exhibits a nearly three fold


drop; the 70 phr compound shows a maximum two fold decreases in the
same temperature range. As expected, the G′ vs. T transition corresponds
to a peak in the tan δ vs. T curve (right graph in Figure 5.35). Below the
transition temperature, changing the CB content has marginal effect, if
any; in the higher temperature region, increasing filler content signifi-
cantly increases tan δ.

5.1.8.4  Effect of Carbon Black Type on G′ and tan δ


The stronger the reinforcing character of a given CB, the larger the elastic
modulus increase. Because the complexity of the aggregate is the most influ-
ential filler characteristic with respect to reinforcement, direct relationships
exist between dynamic moduli and parameters related to the structure of
CB. This well known aspect is illustrated in Figure 5.36, where the plateau
elastic modulus GN0 of 40 phr filled butyl rubber compounds is plotted vs.
the N2 specific area. Low structure blacks, i.e., N990 to N550 marginally affect
the plateau modulus but the reinforcing effect is clearly related to specific
area with higher structure fillers (as clearly seen with ­carbon blacks from the
N200 series).
In fact, the variation in dynamic properties imparted by a given CB is
depending on temperature, as clearly seen when considering tan δ. Figure
5.37 shows the effect of CB type and temperature on tan δ of solution SBR
compounds. Above ordinary temperature (i.e., 20°C), the more complex the
structure of the CB the higher the tan δ, but the maximum tan δ (occurring
at around −20°C, irrespective of filler type or level) tends to decrease as the

10
Butyl rubber, 40 phr furnace carbon black
Plateau elastic modulus G´0N, MPa

N210
6 N242

N219
4
N293
N330
N660
2 N770
N880 N550
N990
0
0 20 40 60 80 100 120 140 160
Specific area N2 , m2/g

Figure 5.36
Effect of carbon black specific area on elastic modulus. (Adapted from A.R. Payne, R.E.
Whittaker. Rubb. Chem. Technol., 44, 440–478, 1971.)
2.0
SSBR Duradene 715 100 Dynamic spectrometer II
N234 Carbon black 50 (Rheometrics Inc.)
1.0 N347 Zinc oxide 3 5% DSA; 10 Hz
Stearic acid 2
N660 Antioxidant 1
Tangent δ vs. tire performance
Polymers and Carbon Black

0.5 Gum Sulphur 1.75


CBS 1.25 Ice Wet Rolling
grip grip resistance
MBT 0.2
0.3

Tangent δ
0.1

Tangent δ
0.05

0.03
–60 –40 –20 0 20 40 60 80 100 120 –80 –60 –40 –20 0 20 40 60 80 100 120
Temperature, °C Reduced temperature at 1 Hz, °C

Figure 5.37
Effect of carbon black type and temperature on damping properties of SSBR compounds; relationship with tire performance.
145
146 Filled Polymers

structure increases. Between −20 and  + 20°C, the higher the structure, the
lower the slope d tanδ/dt, and below −20°C the carbon type (and level) has
no significant effect. The combined effects of CB and temperature on the
dampening properties of a filled rubber compounds are thus, complex and,
as illustrated by the insert in Figure 5.37, it is therefore the choice of the filler
that allows tire engineers to adjust dynamic properties with respect to antag-
onistic requirements. It is easy to understand that ice and wet grip resistance
of a tire are favored by a larger viscous response in dynamic conditions, in
other words by an increased tan δ. The requirements are opposite in what the
rolling resistance is concerned since a dominant elastic response (i.e., lower
tan δ) favors an efficient use of the energy provided by the engine of the auto-
motive vehicle.
In terms of tire performance, the reinforcing effects due to a given grade
of CB are in fact very complex and the relationships between material func-
tions, such as dynamic moduli, and properties of technological significance
are not straightforward. As we said before, many of the relevant properties
for tire application evolve in antagonistic manner when the grade (and/or
the level) of CB is changed. A study by Byers and Patel55 illustrates the dif-
ficulties met by the compounder when trying to improve tire tread perfor-
mances through the selection of an appropriate grade of CB. Figure 5.38 was
drawn using selected data from these authors.
As can be seen, the abrasion resistance improves (lower loss under abra-
sion) as the reinforcing character of the CB increases, and the structure of the
aggregate is the key factor, as demonstrated by the ­variation observed when
considering the N300 family. The heat build-up under repeated flexion and
the rebound evolve in (different) directions with black grade in a manner,
which essentially reflects the increasing tan δ with the reinforcing character
of the filler, both at room temperature (rebound) and at higher temperature
(under repeated flexion). One notes that whatever the technological property
considered, there is a kind of plateau in property gain as the finer CB grades
are considered, i.e., not much gain is observed when moving from N220 to
N110, but the latter black is more difficult to dispersed in a rubber matrix
than the former.

5.1.8.5  Effect of Carbon Black Dispersion on Dynamic Properties


The poorer the dispersion the larger the DSS (Payne effect). Indeed as
shown in Figure 5.39, the elastic modulus drop with increasing strain
amplitude tends to decrease as mixing duration increases. But DSS still
exists for optimally dispersed compound, as may be expected with respect
to curves in the Figure 5.38 which tend to be closer each other as mix-
ing time increases. Such observations are somewhat contradictory with
the simple “filler network” interpretation offered by Payne for the strain
dependent modulus. Indeed, as the dispersion of CB particles improves,
the spatial influence of any expected filler network should widen and the
Abrasion resistance Heat build-up under repeated flexion
20
N326
18 NR/black (50 phr) 120 N220 N110

N347 N234
16
N330 N339
N351
14 N375
N351 100 N330
N347
12 N326
N220

Abrasion loss, g/h


N339
NR/black (50 phr)
Polymers and Carbon Black

Running temperature, °C
N375 N234
10
N110 80
0 20 40 60 80
8
0 20 40 60 80 Reinforcement factor RF
Reinforcement factor RF
Elastic rebound
70
ηcpd
Reinforcement factor : RF = η
gum N351
65 N347
N326 N339
(Mooney) viscosity
SMR L 100 N330 N375
of compound N110
Carbon black 50 (Mooney) viscosity 60

Rebound, %
Zinc oxide 5 N220
of gum rubber N234
Stearic acid 3
BTSa 0.6 55 NR/black (50phr)
Sulphur 2.5

a benzothiazol disulfide 0 20 40 60 80
Reinforcement factor RF

Figure 5.38
Effect of carbon black grade on technological properties relevant in tire tread performances.
147
148 Filled Polymers

SBR/N220 (70 phr)


6

1.5 min
5 2.0 min
2.5 min
Elastic modulus G´, MPa

3.0 min
4
16.0 min

0
0.001 0.01 0.1 1.0
Double strain amplitude
Figure 5.39
Effect of mixing duration on the magnitude of dynamic stress softening.

reverse of Figure 5.39 should be observed. But one has nowadays more
information than at Payne’s time and if there is indeed a network effect,
it concerns both the filler particles and a part of the polymer matrix,
strongly bound through intense particle-polymer interactions, as dis-
cussed hereafter.
It is well known in the rubber industry, particularly in tires manufactur-
ing, that a sizeable reduction in tan δ is obtained through improved micro-
dispersion of fillers, in other terms by loosening-up the filler network that
would result from undislocated agglomerates and/or strongly connected
aggregates. For instance, benefits of extensive mixing have been documented
as well as the positive effects of so-called chemical promoters or coupling
agents, whose role is expectedly to enhance/promote interactions between
CB particles and rubber chains, therefore, to achieve a better dispersion and
hence to reduce hysteresis. Most of those chemicals however either have an
ill documented toxicity potential or are not really in line with present envi-
ronmental concerns.

5.1.9  Origin of Rubber Reinforcement by Carbon Black


It is quite difficult to propose an overall coherent explanation for the rein-
forcing effect of CB and only certain aspects are so far understood. First, the
true origin of rubber–filler interactions is not yet clear and second, as we
Polymers and Carbon Black 149

have previously underlined, the concept of reinforcement is complex since


it concerns a balance of properties, most of them being antagonistic. For
instance, in tire technology, would one want to increase the abrasion resis-
tance of a tread band by changing the grade of CB, it is likely that the wet
skid resistance will be affected. We will describe hereafter some of the most
interesting (or most accepted) theoretical views on the subject.
Payne has proposed a classical splitting of all the effects involved in rubber
reinforcement which can be further refined with respect to contemporary
views (Figure 5.40). In addition to the intrinsic properties of the elastomer
network, the CB particles bring first mere hydrodynamic effects, which are
further enhanced by strong rubber–filler interactions, and interaggregate
interactions which are weaker and depending on strain level.
A number of authors58 have followed the Payne suggestion that the elastic
modulus decrease with increasing strain could be related to the progres-
sive breakdown of an interaggregate network, assumed to exist in a well
dispersed CB filled compound. Accordingly the higher dynamic modulus at
low amplitude would result in a stiff CB network that would be destroyed
as strain amplitude is rising. Such a model is however unable to explain all
aspects of DSS, namely the modulus recovery that is observed when a previ-
ously strained material is left at rest for a sufficient period at the appropriate
temperature.

Weaker
interaggregate interactions
= f(carbon black grade and level )
Elastic modulus

Strong carbon black–rubber interactions


= f(carbon black grade and rubber type )

Hydrodynamic effects
= f(carbon black type and level)

Rubber network properties


= f(rubber type and vulcanization system)

Strain
Figure 5.40
Origin of rubber reinforcement by carbon black particles.
150 Filled Polymers

As we have seen, hydrodynamic effects can somewhat be understood with


respect to the works of Einstein, Guth, Gold, Simha, and others, but most
of the technologically significant effects are due to rubber–CB interactions.
BdR is the most significant evidence for rubber–carbon interactions, which is
readily considered through the effective filler fraction, i.e.:

Φ eff = Φ filler + Φ BdR

This approach was explicitly proposed by Medalia who adapted the Guth,
Gold and Simha equation to modulus (see Equation 5.12), but the effects of
CB on mechanical properties are more or less complicated depending on the
grade and level of CB and, moreover, one cannot expect much from consid-
erations solely based on hydrodynamic effects.
At molecular level, it is fairly obvious that the tightly BdR layer some-
what influences the loosely BdR fraction. There is thus, a reduced segmental
mobility in the neighborhood of a CB particle or, in other terms, a “stiffness
gradient” as one moves away from the surface of CB aggregates, as explicitly
considered by Heinrich and Vilgis59 (see Figure 5.41).
Such a view is very similar to the concept of “mesophase” introduced by
Theocaris for polymer–glass fiber systems60 and other particle-filled com-
posites.61 This concept was developed to explain why the glass transition

Concept of transition layer Concept of mesophase


(Heinrich & Vilgis (1995)) (Theocaris (1987))
z
Mesophase

Fiber
Interface

Matrix
Carbon Bound phase
Extractable
black rubber
rubber Chemisorbed molecules
(irreversible)
Transition layer Physisorbed molecules Constrained
(reversible) matrix
Figure 5.41
Reduced segmental mobility imparted by the proximity of a rigid body with which strong
interactions occur.
Polymers and Carbon Black 151

of composite materials may be reduced in some cases, whereas it may be


increased in others. A mesophase is the region, between the matrix ­material
and the filler particles, both considered as homogeneous and isotropic, whose
thermomechanical properties and fraction are determined from the overall
thermomechanical behavior of the composite.
Involving BdR in theories on reinforcement is surely attractive, with
respect to available experimental data, but because rubber–CB interactions
have essentially a physical nature, they are reversible and, at a given tem-
perature, the actual level of BdR (and hence of rubber–filler interaction) is at
best an equilibrium level between competitive ­adsorption–desorption pro-
cesses. As discussed below several quantitative models for the DSS effects
have been developed.

5.1.10  Dynamic Stress Softening Effect


5.1.10.1  Physical Considerations
One the most typical nonlinear viscoelastic properties of filled rubber
compounds is DSS, i.e., the decrease of dynamic moduli (either the com-
plex G* or the elastic G′ modulus) upon increasing strain amplitude at con-
stant frequency and temperature, a phenomenon commonly referred as the
“Payne effect,” with respect to the extensive studies of A.R. Payne,62–64 whilst
Fletcher and Gent were the first to report it.65 The occurrence of a maxi-
mum in the viscous modulus G′′ is generally associated with the decrease
in G′ but it must be underlined that, if G′ and G′′ are calculated from G* and
tan δ( = G″/G′) measurements, the equality G* = G′ + iG″ is strictly valid in the
linear viscoelastic domain only. However, using a ­mechanical spectrometer
with a highly precise displacement transducer offering a ­resolution of 0.05
μm (IMass Co. Dynastat Mark II), Roland66 reported that, even at amplitudes
for which the modulus is a marked function of strain (γ0 range ≈ 0.001 − 0.01),
an undistorted sinusoidal output is obtained in response to a sinusoidal
input. Despite this stress/strain proportionality (which for unfilled, pure
polymers is typical of the linear viscoelastic domain) the elastic (G′ or E′)
and viscous (G′′ and E′′) moduli exhibit a strong dependence on strain mag-
nitude. This amplitude-dependence under dynamic loads has been observed
with a large variety of elastomers and fillers, and appears thus, a typical non
linear character of filler-loaded vulcanizates. The magnitude of the effect
depends on the temperature, the strain history and other factors. Almost
all the DSS effect is achieved in a first strain sweep and only minor effects
take place in subsequent steps. However, as clearly shown by Wang,67 upon
aging at room temperature the softening effect is recovered, at least partially
in most cases.
Various origins (i.e., physical interpretations) have been assigned to the
DSS effect, all calling to local mechanisms involving interactions either
between filler particles, or between filler particles and the polymer matrix,
152 Filled Polymers

or a combination of both. Not all interpretations lead however to quantitative


workable models. As might be expected, interpretations offered over the time
(so far half a century) somewhat reflect not only the on-going developments in
technical resources and in the accuracy of investigation means, but also the at-
the-time available information regarding the nature of the filler, its structure
and surface quality (i.e., chemistry, physics, and topology), as well as the level
of understanding of the filler–filler and rubber–filler interactions. However,
because very precise measurements have revealed that unfilled materials
exhibit also DSS effects providing the strain magnitude is sufficiently large,
other mechanisms should be considered as well. A most convincing one, in
the case of unfilled vulcanized elastomers is the nonaffinity of the deforma-
tion, which essentially means that cross-links move with respect to a coordi-
nate system embedded in the material, with an obvious entropic character.68
DSS is observed with nearly all filled systems and, accordingly, a number
of mechanisms have been proposed in the literature that we have (somewhat
arbitrarily) sorted out in two main categories:

• Destruction (and reforming) of a so-called “filler network”


• Destruction (and reforming) of a complex “filler–polymer network”

From an engineering point of view, the most useful contributions from the
literature are the ones that lead to quantitative workable models, i.e., math-
ematical relationships that can be used first to fit correctly experimental
data, second to (tentatively) interpret fit parameters in view of the physical or
physico-chemical processes considered. In discussing the physical interpre-
tations of the DSS effect, quantitative workable models are therefore receiv-
ing hereafter privileged considerations.

5.1.10.2 Modeling Dynamic Stress Softening


as a “Filler Network” Effect
The destruction (and reforming) of a so-called “filler network,” assumed to
exist in the material, was initially considered by Payne himself.69 In this model,
adjacent CB aggregates are considered as sufficiently close for their respective
surfaces to interact through interfacial forces of the London–Van der Waals
type.* Such forces would have a significant contribution to modulus at very
low strain but would vanish when the separation between interacting sur-
faces increases. Their contribution to modulus is thus expected to decrease
upon increasing strain in dynamic tests. Filler network breakage is the back-
ground of the well known phenomenological model developed by Kraus70
(see below), in which, under dynamic deformation, physical bonds between

* Van der Waals forces are short distances interaction forces of the order of 4 kJ/mole, to be
compared with hydrophobic interactions (around 8 kJ/mole) and hydrogen bonds (20–40 kJ/
mole, depending of chemical groups involved).
Polymers and Carbon Black 153

(log) G´0

G´´
G´∞
tan δ
G´´

Tan δ

(Log) strain amplitude


Figure 5.42
Dynamic strain softening effect (schematic).

filler particles are continuously broken and reformed. The filler network is
an obvious reference to the percolation theory* and is therefore acceptable
when the filler volume fraction is above the percolation level, in the 12–13%
range for most systems. Because strain dependent dynamic moduli are also
observed when the filler content is below this level, any mechanism based on
the dislocation of a (pure) filler network is therefore not fully satisfactory.
Thanks to progresses in particles observation techniques and to new con-
cepts such “fractal objects” introduced by B. Mendelbrot,71 the filler network
approach was considerably enlightened, namely by taking into consideration
the three-dimensional structure of carbon blacks. The fractal nature of the
filler surface became a key aspect, leading to models that consider aggregates
percolation when the volume fraction of filler is sufficiently high. Huber and
Vilgis72 considered interactions between aggregates through dynamical pro-
cesses of breakage and reformation of the resulting network embedded in
the polymer matrix. When dynamic strain amplitude increases, the perco-
lated network breaks down in smaller and smaller entities.
A number of experimental data have shown that a great variety of filled
rubber vulcanizates exhibit the dynamic strain softening effect schematically
illustrated in Figure 5.42, at given frequency and temperature.† The ­limiting

* It seem that percolation theory has his root in the works of Flory and Stockmayer who used
it to study polymerization processes that may lead to gelation; see D. Staufer, A. Aharony.
Introduction to Percolation Theory. 2nd Ed., Taylor & Francis, London, UK, 1994. ISBN 0 7484
0253 5.
† It is worth noting that dynamic strain softening is observed on filled vulcanizates what-

ever is the strain mode, i.e., shearing, tensile of uniaxial compression. The shear modulus
G* = G′ + iG″ is explicitly considered here, but the same considerations would in principle
apply as well to the tensile modulus E* = E′ + iE″ or the compliance J* = J′−iJ.″
154 Filled Polymers

low strain value of the elastic modulus is G0′ and all experimental data sug-
gest that there is a high-amplitude plateau G∞′ . All quantitative workable
models so far proposed start from this schematic DSS representation.
Apart a few early unsuccessful attempts to assign the effect to the rupture
of a network of (spherical) contacting particles, the first quantitative model
was suggested by G. Kraus,73 with respect to a thixotropic change in the
(filler) microstructure, or in other terms a mechanism of deagglomeration-
reagglomeration of (CB) aggregates. The basic idea of Kraus was that aggre-
gates must be viewed as “soft spheres” such that, over an infinitesimal range,
the interparticle force F increases continuously with the displacement from
an equilibrium distance rm. Interaction between “contacting” aggregates can
therefore, be viewed as a (nonlinear) phenomenon whose characteristic rate
is (∂F/∂r )r =rm. He then referred to an attraction–repulsion mechanism, as suit-
ably modeled with respect to a Lennard-Jones potential, and at the expense
of a few approximations, he established that the difference between the two
limiting elastic modulus values, i.e., (G0′ − G∞′ ) is proportional to N0, the num-
ber of effective interparticle contacts at rest (zero amplitude), the character-
istic rate and the equilibrium distance between soft spheres of radius d/2 to
which aggregates are assimilated (Figure 5.43), i.e., (see details of the devel-
opment in Appendix 5.2):

 ∂F 
(G0′ − G∞′ ) = C1 N 0   (d + rm )2 (5.33)
 ∂r  r = r
m

where C1 is a constant of proportionality.


To account for the strain-dependence of the dynamic moduli, Kraus con-
sidered that, at equilibrium the rates of aggregate deagglomeration and
reagglomeration are equal and assumed that power law functions account
for the strain dependency, i.e.:

kB N γ 0m = kR ( N 0 − N ) γ 0− m (5.34)

where kB and kR are respectively, the rate constant for deagglomeration and
agglomeration (or “breaking” and “reforming” of the filler network), N0 the
number of contacts between aggregates at zero strain, N the number of sur-
viving contacts, γ0 the strain amplitude and m a parameter describing the
strain dependency. By defining a critical strain γc as follows:
2m
k 
γc =  R  (5.35)
 kB 

the following equality is derived from Equation 5.34, i.e.:


N0
N= 2m
(5.36)
γ 
1+  0
 γc 
Polymers and Carbon Black 155

σ rm
10

Potential V(r)
0

ε
rm
–10
0 0.5 1
Interparticle distance r
rm rM

50

∂F
∂r r = r
Force F(r)

m
0
d

–50
0.3 0.4 0.5
Interparticle distance r
Figure 5.43
Kraus model of deagglomeration–reagglomeration of filler aggregates, assimilated to “soft
spheres.”

It is worth making two comments here. First postulating that the same
power law function suits adequately the strain-dependency of the deagglom-
eration and the reagglomeration processes with only a change of sign is a
strong hypothesis because it implies a perfect symmetry of both processes.
Second the definition of the critical strain postulates also (time) symmetry.
Consequently symmetries are being forced in the model equations, as it is
indeed the case (see below).
In agreement with the attraction–repulsion mechanism described above,
the excess modulus of the agglomerate network at any amplitude must be
proportional to N, so

G′ ( γ 0 ) − G∞′ 1 G0′ − G∞′


=    or   G′( γ 0 ) = G∞′ + (5.37)
G0′ − G∞′ γ 
2m
γ 
2m

1+  0 1+  0
 γc   γc 

One sees immediately that γc is the strain amplitude corresponding to


(G0′ +G∞′ )/2 .
156 Filled Polymers

Kraus considered that the deagglomaration and the agglomeration


­ rocesses are deemed to dissipate (strain) energy, in addition to the energy
p
dissipated in straining the vulcanizate. There is thus, an excess loss modulus
that may be taken as proportional to the rate of aggregate deagglomeration
(or reagglomeration), i.e.:

G′′( γ 0 ) − G∞′′= c kB N γ m0

where c is a constant of proportionality. Substituting for N (Equation 5.36)


and using Equation 5.33 leads to:

C γ 0m (G0′ − G∞′ )
G′′ ( γ 0 ) − G∞′′= 2m
(5.38)
γ 
1+  0
 γc 

The function G″(γ0) has a maximum at γ0 = γc so that:

m m
γ  γ 
2 0 2 (Gm′′ − G∞′′)  0 
G′′ ( γ 0 ) − G∞′′  γc   γc 
=    or   G′′ ( γ 0 ) = G∞′′+ (5.39)
(Gm′′ − G∞′′) γ 
2m
γ 
2m

1+  0 1+  0
 γc   γc 

where Gm′′ is the maximum loss modulus at γ0 = γc.


The relationship for the loss tangent is then easily obtained, i.e.:

2
 γ  m/2  γ 0  − ( m/2 ) 
G∞′′  0  −   + 2 Gm′′
 γ c   γc  
tan δ = −m m
(5.40)
 γ0   γ0 
G0′   + G∞′  
 γc   γc 

The above equations require six parameters, G0′ , G∞′ , m, γc, Gm′′, and G∞′′ .
Good estimates of G0′ , m, γc, and Gm′′ can generally be directly extracted from
experimental data sets but both the elastic and the viscous moduli at infinite
strain are at best extrapolated values. Two parameters, m and γc, are common
to both G′ ( γ 0 ) and G′′( γ 0 ) . Several authors have applied the Kraus equations
to a number of experimental data with fairly good correlation, at least for
G′(γ0), but viscous modulus data at very low strain (i.e., < 10−3) are not well
accounted for by Equation 5.39. For instance, Ulmer74 tested the Kraus rela-
tionships with data on various CB filled compounds, either his own, or from
literature. Whatever the testing mode (i.e., shear, compression or torsion), non-
linear fitting of Equation 5.37 on G′ (or E′) data gave correlation coefficient r2 in
the 0.99 +  range, at least for CB loadings lower than 60 phr. Fitted values for m
Polymers and Carbon Black 157

8 Gínf G´m γc m r2
N110 0.659 1.026 1.428 0.547 0.692
Elastic modulus, MPa

6 N330 0.477 0.713 1.808 0.706 0.640


N660 0.224 0.300 2.174 0.491 0.709
4
1.2

Viscous modulus, MPa


2
0.8
0
0.01 0.1 1 10 100
Strain, %
0.4
Gínf G´0 γc m r2
N110 2.621 6.209 1.376 0.614 1.000
N330 2.363 4.581 1.471 0.573 0.999 0
0.01 0.1 1 10 100
N660 1.825 2.414 1.761 0.344 0.998 Strain, %
Figure 5.44
Kraus model to fit experimental G′ and G′′ data. (Data from M. Gerspacher, C.P. O’Farrell,
C. Tricot, L. Nikiel, H.A. Yanga, ACS Rubb. Div; Mtg, Louisville, KY, Oct. 8–11, 1996. Paper 74
on SBR/carbon black (60 phr) compounds.)

were in the 0.5–0.7 range for G′(γ0), in line with m = 0.6 as considered by Kraus
when probing his model, but significantly different m values were found for
G′′ data. The critical strain γc was found to decrease with increasing CB load-
ing. To further document how the Kraus model meets experimental data, we
used a nonlinear algorithm (i.e., Marquart–Levensberg) to fit data obtained
by Gerspacher et al.75 on a series of SBR filled compounds (see Appendix 5.2),
with Equations 5.37 and 5.39 (Figure 5.44).
As can be seen, the model fits reasonably well the experimental G’ data
but cannot meet the asymmetric shape of G” vs. strain data. The exponent m
is depending on the reinforcing character of the filler, as well as the ­critical
strain γc. A similar assessment of the Kraus equations with a silica (40 phr)
filled polydimethylsiloxane composite76 yielded the same conclusions regard-
ing the low strain deficiencies of Equation 5.39. For such a silica filled mate-
rial, fit parameter values were as expected different, i.e., γc ≈ 0.02 and m ≈ 0.45
for G′(γ0) ; γc ≈ 0.01 and m ≈ 0.35 for G″(γ0).
As previously commented, the postulate considered by Kraus (i.e.,
Equation 5.34) imparts symmetries in both the G′(γ0) and the G″(γ0) functions.
If indeed, experimental data support an horizontal symmetry for the elastic
modulus, with respect to the mid modulus value, no vertical symmetry (with
respect to γc) is generally observed for the viscous modulus. The deficien-
cies of the Kraus model are therefore, embedded in the starting postulate.
Various modifications have been proposed to account for the nonsymmetri-
cal behavior of G′′, without changing the physical ideas leading to the model.
Using different strain exponents for the deagglomeration and reagglomera-
tion processes (Equation 5.34) was probed by Ulmer who concluded that it
158 Filled Polymers

was not a successful alternative to the Kraus approach. He showed however


that adding an (empirical) exponential term, not associated with the filler
network but reflecting an additional (exponential) decrease of the viscous
modulus upon increasing strain, improved considerably the G″(γ0) descrip-
tion. He then suggested rewriting Equation 5.39 as follows:
m
γ 
2 (Gm′′ − G∞′′)  0 
 γc   γ 
G′′ ( γ 0 ) = G∞′′+ + Gk′′ exp  − 0  (5.41)
γ 
2m
 γk 
1+  0
 γc 

where Gk′′ is a drop in viscous modulus assumed to be proportional to N0


the number of contacts between aggregates at zero strain (i.e., Gk′′ = Ck N 0 ; Ck
being a constant) and γk another critical strain. Values for Gk′′ and γk must be
obtained through nonlinear curve fitting, which requires obviously excellent
quality data in the very low strain range but also suitable initial guess values
for the non linear regression algorithm. As pointed by Ulmer, the ratio of
the exponential term to the second right term in Equation 5.41 is less than
0.001 in most cases. Therefore, Gk′′ ≈ Gm′′ − G0′′ is a good initial guess value
and, whilst somewhat depending on the polymer and the CB loading, the
­corresponding γk guessed value can be taken as considerably smaller than γc.
By comparison with the right graph in Figure 5.44, Figure 5.45 ­demonstrates
the benefit in using Equation 5.41 to fit G′′ data from Gerspacher et al. (see
Appendix 5.3 for details on the nonlinear fitting process).
To consider that deagglomeration and agglomeration are two symmetri-
cal processes is surely a strong hypothesis and likely the reason for the
deficiencies of the Kraus model. In order to somewhat circumvent it, recent
theoretical developments were made with an explicit reference to the fractal
description of CB aggregates, and by considering that, above a percolation
threshold, highly branched aggregates can flocculate and form a secondary

1.2
G´´∞ G´´m γc m
Viscous modulus, MPa

N110 0.120 1.036 1.295 0.519


0.8 N330 0.087 0.723 1.351 0.466
N660 (–0.28) 0.300 1.271 0.162

G´´k γk r2
0.4
N110 0.572 0.199 1.000
N330 0.371 0.156 0.999
N660 0.130 0.042 0.985
0
0.01 0.1 1 10 100
Strain, %
Figure 5.45
Kraus model as modified by Ulmer on SBR/carbon black (60 phr) compounds.
Polymers and Carbon Black 159

network. A  kinematics is associated with such an aggregates flocculation


­ rocess, based on the hypothesis that complex tri-dimensional entities like
p
CB aggregates are not fixed in the space of the material but can perform
random motions such that they fluctuate around their mean position in the
rubber matrix.77–80 (Note that Klüppel et al. refer to their approach as the
“cluster–cluster-­aggregation” (CCA) model; in this book, we use the term
“cluster” to describe the large scale structures that silica “string of pearls”
develop through hydrogen bonding; CB aggregates entangle through other
mechanisms where chemical bonding plays no role; therefore we renamed
their model aggregate floculation for consistency.)
By considering that, in their working temperature window, elastomers are
by definition amorphous polymers on their rubbery plateau and beyond,
the hypothesis that aggregates do entangle (or flocculate) through random
motions in the rubber matrix is of course reasonable. In the author’s opinion,
one of the key aspects of the aggregates entanglement model is the percolation
threshold, i.e., the filler volume fraction Φ* below which the stress applied on
the filled compound is essentially supported by the polymer matrix around
the aggregates that are sufficiently separated from each other to practically
not interfere at all. Above Φ*, entangled aggregates form a network which
has the capability to support and transmit stress, in such a manner that it is
necessary to consider the modulus of this tenuous secondary filler structure
(Figure 5.46).
In the rubber matrix, flocculated (or entangled) aggregates form a kind
of “filler network” whose elastic modulus at small strain amplitude would
depend on their fractal connectivity. The low strain elastic modulus of the
filler network can be evaluated by considering that entangled aggregates
may store energy when strained, through bending–twisting deformation of
the inter-aggregates connecting points. An interpretation of the Klüppel et al.
proposal, supported by some numerical assessment (see Appendix 5.4) leads
to the following practical relationships. The aggregate solid fraction is firstly
considered with respect to a fractal description, i.e., (see Figure 5.46, left):

3− F
 d
Φ A (D) =   (5.42)
 D

where D and d are the diameters of the aggregate and the elementary particle
respectively, and F the fractal dimension of the aggregate. Then, with respect
to proposals for flocculated materials, the elastic modulus of an aggregate GA
is approximated as follows:

( 3 + FB )/( 3 − F )
GA = Gp ⋅ Φ A (5.43)

where FB is the fractal dimension of the aggregate backbone and Gp an


averaged elastic modulus of the aggregate with respect to all its possible
160

D a

Figure 5.46
Aggregates flocculation process according to Klüppel et al.
Filled Polymers
Polymers and Carbon Black 161

angular deformation. In other terms, Equation 5.43 describes the small


strain ­modulus of the flocculated aggregates as the product of a local mod-
ulus for the ­elastic deformation of aggregate–aggregate connecting points
and a geometrical factor, based on the fractal description of the aggregate. In
fact with respect to the BdR fraction, it is necessary to consider that there is a
layer of immobilized polymer at the (available) surface of the aggregate. For
CB aggregates, the typical fractal dimensions are respectively, F ≈ 1.8 and
FB ≈ 1.2, so that GA varies with the power ~3.5 of the aggregate solid frac-
tion. The CB concentration dependence of the small-strain dynamic ­elastic
modulus G′ above 12–13% filler, as reported by several authors for various
rubber compounds (i.e., log G′ ∝ 3.5 · log Φblack) is therefore considered as
a strong experimental support for this fractal approach. The mechanically
effective solid fraction of the aggregate, i.e., ΦefA(D), must include the tightly
BdR. A bound layer of thickness a on the exposed surface of the aggregates
can be considered but, because by nature an aggregate is a random object,
a precise estimation is impossible. The following approximation gives how-
ever reasonable results when running basic numerical assessment, i.e., (see
Appendix 5.4):

π 2π 2  d  
N p (D) ⋅  ( d + 2 a ) −
3
a  3  + a − a  
 6 3  2   ⋅
Φ efA (D) = (5.44)
π 3
D
6

where Np(D) is the number of elementary particles in an aggregate of diame-


ter D. The term [(π/6)(d + 2a)3] is obviously the surface of an elementary sphere
of diameter d + 2a and the term

2π 2  d  
a  3  + a − a 
3  2   

takes into account the inaccessible surface at the contact point between
neighboring particles. Using typical values for N330 CB, i.e., Np(D)  = 200
particles, d = 30 nm, D = 200 nm, and a = 2 nm, the first term yields
20,580 nm3 and the second 410 nm3. This means obviously that the frac-
tal description of CB suits well “open” branched aggregates. Because the
tightly BdR layer is relatively thin ( i.e., a << d), a good approximation of
Equation 5.44 is:

N p (D) ⋅ [(d + 2 a)3 − 6 d a 2 ]


Φ efA (D) = ⋅ (5.45)
D3

With respect to Equation 5.37, the aggregates flocculation model allows


thus, to consider that the low strain modulus of a CB filler compounds
162 Filled Polymers

(whose filler level is above the concentration threshold Φ*) can be estimated
from:
( 3 + FB )/( 3 − F )
 N p (D) ⋅ ( d + 2 a )3 − 6 d a 2  
G(Φ) = G(0) + Gp ⋅    ⋅ Φ (5.46)
 D3 
 

where G(0) is the modulus of the elastomer and Gp the averaged elastic modu-
lus of the aggregate. Most raw rubbers have low strain modulus of the order of
a few MPa, but to offer a reasonable estimate of Gp is obviously quite difficult.
Klüppel et al.78 suggested to consider 10 GPa, with respect to the modulus
of para-crystalline carbon, but in such a case, Equation 5.46 yields generally
excessive values. If for the sake of evaluating the pertinence of the aggregate
flocculation model, one considers that between the modulus of a raw rubber
and the elastic modulus of the aggregate there is a factor of 1000, then the
graph given in Figure 5.47 can be drawn, in comparison with the well known
Guth and Gold equation.
As can be seen, below the percolation level (around 0.13), the aggregate
flocculation model is not relevant, as well as the lower predicted modulus,
in comparison with the Guth and Gold approach. For higher filler volume
­fraction, Equation 5.46 predicts a larger increase of the compound’s modulus
as filler volume fraction increases, in better agreement with usual experi-
mental data on compounds with reinforcing carbon blacks than the Guth
and Gold model. Figure 5.47 supports the comment that, whilst perfectible,
the kinetic flocculation of aggregates with a fractal description of the latter
might be the right theoretical concept for (CB) filled polymers.

10
Klüppel & et al.
Compound modulus, MPa

Guth & Gold, 8% BdR


Guth & Gold, 30% BdR

0
0 0.05 0.10 0.15 0.20 0.25
Filler volume fraction Φ
Figure 5.47
Comparing the aggregate flocculation model (Equation 5.46) with the Guth and Gold equation;
data used : Np(D) = 200 particles, d = 30 nm, D = 200 nm and a = 2 nm, G(0) = 1 MPa, Gp = 1 GPa; the
Guth and Gold equation is used by considering either 8% or 30% bound rubber.
Polymers and Carbon Black 163

Flocculated aggregates might be the main source of enhanced compound


modulus with higher CB volume fraction but it is quite obvious that as strain
increases, such a tenuous secondary structure will be destroyed and will
not participate anymore in transmitting the stress through the material.
Connecting points in the aggregate flocculation model can be assimilated to
the interparticle contacts between soft spheres in the Kraus model and a rate
equilibrium between dislocated and flocculated aggregates be considered
as well. Heinrich and Klüppel79 used this idea to derive an equation for the
DSS, as follows:
−τ
  γ 2m
G′( γ 0 ) = G∞′ + (G0′ − G∞′ ) ⋅ 1 +  0   (5.47)
  γ c  

where τ ≈ 3.6 is the elasticity exponent of percolation, i.e., (3 + FB)/(3 − F), and


the only difference with Equation 5.37 derived by Kraus. Unfortunately,
no equation is given for the corresponding variation of G′′ with strain
­amplitude. A numerical comparison of Equations 5.37 and 5.47 shows that
no much is gained in modeling the DDS with respect to the aggregate floc-
culation approach (see Appendix 5.4). For equal initial G0′ and final G∞′
moduli, superimposed curves are obtained with substantial differences in
models’ parameters. Equation 5.47 still exhibits and horizontal symmetry (as
expected since the same rate equilibrium is considered), the upper and lower
limits of the term [1 + (γ0/γc)2m]−τ are still 1 and 0 and, at the critical strain γc,
this term obviously reduces to 2−τ (≈0.0825 with τ = 3.6), instead of 2 with the
Kraus model. With the latter, the critical strain corresponds to the mid modu-
lus value, with Equation 5.47, the critical strain corresponds to a modulus
equal to G0′ ⋅ 2 − τ + G∞′ ⋅ ( 1 − 2 − τ ) .
In the models described some far, very little consideration was brought
to the nature of the matrix in which CB particles are dispersed, except the
Ulmer’s modification of the Kraus model, which introduces in the G″(γ)
model an additional term, explicitly not associated with the filler network,
but reflecting the viscous modulus variation of the polymer upon increasing
strain. As we have seen, the Ulmer proposal leads to an excellent description
of the asymmetric variation of the viscous modulus with strain amplitude.
In the author’s knowledge, van de Walle et al.81 were the first to objectively
assign a role to the polymer matrix by considering that the complex moduli
of filler rubber compounds result from the superimposition of a linear vis-
coelastic behavior of the polymer and a nonlinear effect of the aggregate
when their contact situation changes upon increasing strain. Firstly they
focused on the likely microscopic interaction between two aggregates, by
considering how Van der Walls forces might evolve as the particles are dis-
placed from their equilibrium positions (i.e., at zero strain). Such a starting
point is not much differing from the Kraus views when he assimilated CB
aggregates to interacting “soft spheres,” but van de Wall et al. give explicitly
164 Filled Polymers

a (viscoelastic) role to the polymer matrix. Indeed they consider that upon
strain both aggregates of a pair are moved from their respective equilib-
rium positions, in such manner that a strain is exerted between them, owing
to polymer–filler interaction. This strain can be modeled with respect to a
spring-and-dashpot system. With the assumption that the relative motion
of both aggregates is sufficiently slow for the process to be quasi-static (i.e.,
near a stationary state), the interaction aggregate-polymer-aggregate can be
modeled with a set a very simple equations. As in the Kraus approach, it is
considered that, under cyclic deformation, both aggregates are continuously
set apart and brought close, but with an increase in the energy loss owing to
friction between polymer chains during the separation-and-reformation of
the pair of aggregates. Because a spring-and-dashpot model is considered,
the polymer role is explicitly viscoelastic nonlinear.
A force F exists between both aggregates of the pair, which varies with the
strain γ0 and such considerations lead in fact to an idealized model whose
microscopic complex modulus has the two standard elastic and viscous com-
ponents, i.e., g* = g′ + ig″. Therefore, an energy loss E1 is associated with each
hysteretic cycle, such that:

E1 ∝ g 0 γ b ∆γ (5.48)

where g0 = dF/dγ0, γb the strain below which there is no energy loss because
the hysteresis does not take place and ∆γ the amplitude of the hysteretic cycle.
Using the analogy E1 ∝ G′′ γ 20 , one obtains:

0 if γ0 ≤ γb
g ′′( γ 0 , γ b ) = E1 (5.49)
if γ0 > γb
γ 20

By considering a linear variation of F vs. γ0 in the interval [0, γ0], an esti-


mate is obtained for the effective g′, which leads to:
3
γ 
g ′( γ 0 , γ b ) ∝ g 0  b  (5.50)
 γ0 

It follows that the microscopic viscoelastic function g*(γ0) is given by:

g0 if γ0 ≤ γb
g* = γ 
3
∆γ  γ b 
2
(5.51)
g0  b  + i g0 if γ0 > γb
 γ0  γ b  γ 0 

The behavior of the overall system is considered as the sum of the individ-
ual contributions of an infinite number of pair of aggregates, each having
of course a different γb. One must therefore, introduce a weighing function
Polymers and Carbon Black 165

N(γb)d(γ0) which gives the number of pair of aggregates that are split when
the strain increases from γb to γb + dγ0. The complex modulus of the macro-
scopic system is eventually obtained as:
∞ ∞ 3 ∞ 2
 γb   γb 
*


G ( γ 0 ) = G + W ( γ 0 ) dγ 0 +
*

0

0
 γ  W ( γ 0 ) dγ 0 + i h
0

0
 γ  W ( γ 0 ) dγ 0 (5.52)
0

where G∞* is the complex modulus of the rubber matrix, W(γ0) = g0(γ0)N(γ0)


an overall weighing function and h the average ∆γ/γb. On the right side of
Equation 5.52, the second and the third terms account for the real (elastic)
part of the complex modulus, and the last term for the imaginary (viscous)
part. According to the model, only W(γ0) is material dependent and there is a
relation between G′ and G′′ since they are both derived from the convolution
of the same function W(γ0).
The practical use of this so-called van de Walle, Tricot and Gerspacher
(VTG) model is not straightforward and was somewhat described in fur-
ther publications by the authors.75,82 The difficulty is to properly evaluate the
weighing function W(γ0) from experimental data. It is in fact more convenient
to calculate it from G″(γ) data, using the differential form of the viscous term
in Equation 5.52, i.e.:

1  dG′′( γ 0 ) 2 
W (γ 0 ) = + G′′( γ 0 )  (5.53)
h  dγ 0 γ0 
and the calculation procedure is somewhat explained in the original publi-
cations.* The capabilities of the above model in meeting experimental G′(γ)
data have been documented by the authors, but unfortunately such a dem-
onstration is not offered for G″(γ). With respect to the mathematical form of
the last right term in Equation 5.52, one would expect a vertical symmetry for
G″(γ), in contradiction with most experimental observation. Apart the merit
of introducing the viscoelastic nature of the polymer matrix in the modeling
of the DSS effects through the “filler network” approach, the VTG model is
not offering real advantages over the Kraus approach.
In order to circumvent some deficiencies of the Kraus model, namely the
fact that the whole set of model parameters has to be reconsidered if the
frequency is changed, Lion et al.83,84 proposed a interesting phenomeno-
logical theory which leads also to a six-parameter model for the DSS effect.
Both the frequency and the amplitude are taken into account with this
model and by interpreting the observed history and recovery effects on
the elastic and viscous moduli as manifestations of thixotropy, a so-called

* Note that in the referred publications, it seems that there are a number of misprints, likely
due to the easy confusion between γb, the critical strain for aggregate–aggregate interaction
and γ0, the strain amplitude. Equations 5.52 and 5.53 have been rewritten by the author in
agreement with his understanding of the modeling approach.
166 Filled Polymers

“intrinsic time” z(t) is introduced, instead of the physical time t. This


­i ntrinsic time is expected to account for the process-dependent relaxation
times of any thixotropic material. In this respect, the Lion et al. approach
considers also that DSS is reflecting a filler network effect, whilst it is not
stated in such terms by these authors. The theory is formulated in the time
domain and therefore can relate the stress to any strain history. Both the
elastic and viscous moduli depend on a “reduced frequency” that takes
into account the strain amplitude and the angular frequency. It follows
that, when the amplitude changes, the moduli curves are not only shifted
but also deformed. In one of the latest versions of this theory, 84 a fractional
linear model is used that consists of a spring of modulus E1 in parallel with
another spring of modulus E2 in series with a “fractional” dashpot of viscos-
ity ηβ = (E2 ⋅ λ)β, with λ a characteristic time and β the fractional exponent.
The ­resulting constitutive equations are then formulated with respect to
the intrinsic time z(t) of the material, whose current state of microstructure
is expressed through an internal variable q(t) that determines the temporal
behavior of the intrinsic time z(t).
Through quite a rigorous treatment that complies with both the mathemati-
cal foundations of fractional calculus and the second law of thermodynamics,
the following practical relationships are obtained for the strain dependence
of the elastic and viscous moduli:

 ω λ  2 β  ω λ  β  π 
E2   +  cos  β  
 a ( γ , ω )   a(γ , ω )  2

G ′( γ , ω ) = E1 + 2β β
(5.54a)
 ωλ   ωλ   π
1+   + 2  cos  β 
 a(γ , ω)  a(γ , ω)  2

β
 ωλ   π
E2  sin  β 
 a( γ , ω )   2
G′′( γ , ω ) = 2β β
(5.54b)
 ωλ   ωλ   π
1+  + 2 cos  β 
 a( γ , ω )   a( γ , ω )   2

where a(γ,ω) is the intrinsic time depending on both the strain and the
­frequency, as follows:
2
a( γ , ω ) = 1 + b γ (ω τ)α (5.54c)
π

In the above set of equations, E1 and E2 are elastic (spring) moduli, ω the
frequency, λ = η/E2 a characteristic time (with η the dashpot’s fractional vis-
cosity), β the viscosity fractional exponent, b a proportionality constant, τ and
α respectively, a characteristic time (≈1 s) and another fractional exponent.
It is quite clear from Equation 5.54c that the intrinsic time varies linearly
Polymers and Carbon Black 167

40 8
E1 + E2
ηβ
E1

Viscous modulus G´´, MPa


G´ḿax
Elastic modulus G´, MPa

30 6
E2
2E1 + E2
20 2 4

10 γc 2
γc
E1

0 0
10–4 10–3 10–2 10–1 100 101 10–4 10–3 10–2 10–1 100 101
Strain amplitude Strain amplitude
Figure 5.48
Dynamic strain softening as modeled by Lion et al.84; Equations 5.54a and 5.54b were used in
drawing the curves with the following set of parameters: E1 = 6 MPa, E2 = 31 MPa, λ = 419.36 s,
β = 0.495, ω = 10 Hz, b = 51000, τ = 1 s, α = 0.44.

with the strain amplitude and nonlinearly with the frequency (providing
the exponent α is not equal to 1). Figure 5.48 illustrates some basic features
of the model, by showing that (E1 + E2) and E1 are respectively, the upper and
lower limits for the elastic modulus and that the maximum viscous modulus
occurs at a critical strain given by:

π (ω λ − 1)
γ c (ω ) = (5.55)
2 b (ω τ)α

At the critical strain, the elastic modulus is a mid-modulus value (equal to


(2E1 + E2)/2 and thus not depending on frequency). The model predicts that
as the frequency increases, both the G′ and G′′ curves are shifted along the
strain scale toward higher strain values.
From a theoretical point of view, the Lion et al. model has the merit to
approach the DSS effect by applying constitutive laws formulated on the basis
of fractional calculus, in other terms by formulating the behavior of materials
with respect to fractional time derivatives of stress and strain; an approach
that in principle requires only a small number of material constants to express
the material properties in the time or the ­frequency domain. However, deriv-
ing model parameters from experimental data is not straightforward and, for
instance Lion et al. had to use a ­stochastic Monte Carlo method to estimate
the model parameters for a comparison with experimental data on 60 phr
CB filled rubber compound.84 Moreover, mathematical handling of the above
equations (see Appendix 5.5) shows that, like the Kraus model, this one exhib-
its also horizontal symmetry for the G′ curve and vertical symmetry for the
G′′ curve, and is therefore not expected to perfectly meet experimental data, at
least in its present state of development.
168 Filled Polymers

Modeling DSS as a “filler network” effect is, in the author’s opinion, still
far from completion and except the empirical modification introduced by
Ulmer, no one of the proposed theories gives satisfactory results in meeting
both the experimental G′(γ) and G″(γ) functions. So far collected data on quite
a large number of (CB) filled systems demonstrate the universality of the
effect and to be fully satisfactory any theory should meet experimental data
in all aspects. The major role played by filler particles interacting with each
other to control the low strain modulus is surely well demonstrated by the
excellent fitting of G′(γ) data by the models described above, with obviously
some approaches being (slightly) better than others. The recurrent failure
of all the models in meeting the asymmetric shape of experimental G″(γ)
curves suggests that any reasoning simply based on an equilibrium between
deagglomeration and reagglomeration of (CB) aggregates, whilst correct in
what energy storage processes (i.e., elastic) are concerned, is necessarily lim-
ited when dissipative phenomena are concerned. In addition to the filler net-
work effect, there is obviously a contribution, which is not simply that of the
pure polymer, but has a dissipative nature and is larger at high strain than
at low strain. The successful Ulmer’s proposal in well fitting G″(γ) deserves
consideration and calls for further theoretical developments.

5.1.10.3 Modeling Dynamic Stress Softening as


a “Filler–Polymer Network” Effect
The likely role in the DSS effect of a complex soft network, involving filler
particles and a part of the polymer matrix, finds its root in the measure-
ment of BdR, otherwise identified as an important aspect of the properties
of uncured filled compounds. A number of authors have developed argu-
ments favoring the filler–polymer network approach and have reported
convincing experimental evidences, but only a few went to workable quan-
titative models. Disentanglement of the bulk polymer from the rubber frac-
tion bounded to the filler surface was discussed by Funt85 as a contribution
to the effective crosslink density of the vulcanizate. Between the bulk rub-
ber and the tightly BdR shell, there is a transition zone (i.e., the loosely
BdR) whose level of entanglement is expected to decrease with increas-
ing strain amplitude. At high strain, when many of the entanglements in
the loosely BdR region have been disengaged, the modulus is essentially
depending on the crosslink density and less (or not) on the type of filler
used. The chain softening of a glassy polymer shell surrounding filler
particles as recently discussed by Montes et al.86 exploits a similar idea. It
must be noted that such approaches consist in fact in considering locally a
Mullins effect.
With a reference to the well known Langmuir’s theory87 for the equilib-
rium adsorption–desorption of gas molecules on solid surfaces, Maier and
Göritz88 developed a quantitative model by considering the kinematical
aspects of an adsorption–desorption mechanism of rubber chain segments
Polymers and Carbon Black 169

1 Local area on
carbon black particle

Site occupied by
a stable “link”

Isolated site, available


2
for an unstable “link”

Free site

3 4
Rubber chain
Segments
Figure 5.49
Interactions between rubber segments and carbon black surface according to Maier and
Göritz.

on appropriate sites on the filler surface. Stable (strong) and unstable (weak)
“bonds” between polymer segments and (hypothetical) interactive sites on
the filler particles are expected to occur depending on the strain amplitude.
As illustrated in Figure 5.49, these authors describe the surface of a CB aggre-
gate in a rubber compound as locally consisting of either free sites, or sites
occupied by stable “links” or isolated sites occupied by unstable “links.”
Rubber–filler interactions sites are like “knots” whose effect superimposes
to chemical reticulation. With reference to the theory for rubbery ­elasticity,
in a filled network, the dynamic modulus is proportional to the overall
networking density N and the temperature T, according to (see details in
Appendix 5.6):
G′ = N kB T (5.56a)

where kB is the Boltzmann constant and with three contributions for N:

N = N chem + N stable + N unstable (5.56b)

Nchem is the fraction of the overall networking density due to vulcaniza-


tion knots, Nstable and Nunstable the fraction due to stable and unstable rubber–­
particle “links.” The adsorption–desorption equilibrium is considered
through the following equality:

Φ free × vads = Φ occ × vdes (5.57)

where Φfree and Φocc are the fractions of free and occupied sites, respec-
tively on CB particles, vads and vdes, the rates of asdorption and desorption.
170 Filled Polymers

The adsorption rate is assumed to be constant, while vdes is proportional to


strain γ, i.e., vdes = Kγ, where K is a constant.
It follows that:
1 1
Φ occ = = (5.58)
Kγ 1 + cγ
1+
vads

The fraction of unstable links, which depends on the strain, is related to


the fraction of free sites times the number of isolated sites per volume unit,
i.e.: Nunstable(γ) = Φocc(γ)  ×  Nisolated, and the following relationship for the elastic
modulus is easily derived:
1
G′( γ ) = Gstable
′ + Gunstable
′ × (5.59a)
1 + cγ


where Gstable = ( N chim + N stable ) kB T and Gunstable
′ = N isolated kB T. A similar rea-
soning leads to the relationship for the viscous modulus, i.e.:

G′′ ( γ ) = Gstable
′′ + Gunstable
′′ × (5.59b)
(1 + cγ )2

In order to assess how the Maier and Göritz model meets experimental
data, we used a nonlinear algorithm (i.e., Marquart–Levensberg) to fit data
obtained by Gerspacher et al.75 on a series of SBR filled compounds (see
Appendix 5.6), with Equations 5.59a and b (Figure 5.50).

8 G´śtable G´únstable c r2
SBR/60 phr black cpds
N110 0.661 1.507 1.243 0.970
Elastic modulus, MPa

6 N330 0.492 0.903 1.027 0.925


N660 0.253 0.191 0.795 0.920
4 1.2
Viscous modulus, MPa

2
0.8
0
0.01 0.1 1 10 100
Strain, %
0.4
G´stable G´unstable c r2
N110 2.328 4.072 0.679 0.997
N330 2.236 2.421 0.642 0.999 0
0.01 0.1 1 10 100
N660 1.917 0.445 0.708 0.999 Strain, %
Figure 5.50
Maier and Göritz model to fit experimental G′ and G′′ data. (Data from M. Gerspacher,
C.P. O’Farrell, C. Tricot, L. Nikiel, H.A. Yanga, ACS Rubb. Div; Mtg, Louisville, KY, Oct. 8–11,
1996. Paper 74 on SBR/carbon black (60 phr) compounds.)
Polymers and Carbon Black 171

As shown in Figure 5.50, Equation 5.59a meets reasonably well the dynamic
strain softening of G′; however Equation 5.59b cannot meet the asymmetric
shape of experimental G′′ vs. strain curves.
As can be seen, the dynamic strain softening of the elastic modulus G′
drop is indeed well fit by the model, which supports the view that the elas-
tic modulus decreases with higher strain because there are less and less
unstable links. The maximum in G′′ vs. γ is well taken into account by the
model, as could have been expected with respect to the quadratic term in
Equation 5.59b but the asymmetric shape of the experimental curves is not
at all met by the model. In line with the reasoning by Maier and Göritz,
the quantities related to the unstable part of the rubber–filler network, i.e.,
G′unstable and G′′unstable significantly decrease with the decreasing reinforcing
character of the CB. However, for a given compound, the fit parameter c is
clearly different for the G′ and the G′′ curves, in contrast with the model’s
expectation.
When compared to filler network breakdown models for DSS, the proposal
by Maier and Göritz attributes a key role to (transient) interactions between
chain segments of the rubber network and the filler surface, by splitting the
adsorbed chains between stable and unstable “links.” No indication is given
to the exact nature of the “links,” except that some (if not all) are not perma-
nent; a view that is well compatible with a number of information about BdR.
As for the Kraus model, the postulate and assumptions considered by Maier
and Göritz inevitably lead to symmetries in the mathematical form of both
the G′(γ0) and the G″(γ0) functions. Again the horizontal symmetry for the
elastic modulus is generally well met by experimental data but no vertical
symmetry is generally observed for the viscous modulus. The most critical
assumption in this respect is the equilibrium between adsorption and des-
orption of chain segments on filler particle sites.
Whilst likely restricted to the case of silica filled systems, a further refine-
ment was brought to this model by Ladouce–Stelandre et al.89 who considered
that the mechanism was thermomechanically activated, with a dependence
on the amount of “free sites” available at the (silica) filler surface. But, of
course in the case of silica filled systems, the interaction sites are well iden-
tified, i.e., the silanol groups with their specific chemistry, as we shall see
hereafter.
The response of a complex soft network involving filler particles and the
BdR was also considered by Wang.90 A soft network results from contacts
between particles through elastomer layers which, in a glassy state near the
filler surface, exhibit a gradually decreasing modulus with the ­distance from
the surface. In addition, occluded rubber likely entrapped in filler–rubber
clusters behaves mechanically as a filler. It is interesting to remark that the
concept of a pseudo-glassy shell of polymer surrounding filler particles and
whose rigidity is gradually decreasing as one moves away from the particle
surface, is very similar to the basic ideas for the so-called shear lag model,
developed for short fiber reinforced systems, as we will see in Chapter 7.
172 Filled Polymers

It must be noted that, apart the initial filler–filler network considerations


of Payne and Kraus, the subsequent models essentially recognized the
­fact—today widely accepted—that, except in very highly filled systems,
direct filler–filler contacts are very unlikely. Explicitly stated by Aranguren
et al.,91 in the case of silica-silicone systems however, it must be noted that
successful BdR measurements imply that filler particles surface is com-
pletely wetted by the polymer. Therefore, contacts between filler aggregates
can occur only through the polymer. The differences in the various models
arise either from the description of the filled system or from the manner the
local thermodynamics is treated.

5.2  Thermoplastics and Carbon Black


5.2.1 Generalities
Contrary to rubber materials, CB is not used in thermoplastics for reinforce-
ment purposes. There are essentially three reasons for CB addition to a
thermoplastic:

• Tinting or pigmentation (in gray or black)


• Reducing UV sensitivity
• Modifying the electrical properties (conductivity)

The two first applications involve only a few percentage weight of CB,
largely below any percolation level above which significant changes would
occur in rheological and/or mechanical properties. The latter application
implies obviously that a “threshold limit,” evidently related to the percola-
tion level, is reached through CB loading, in order to have, in the thermo-
plastic matrix, a network of particles to act as an internal electrical conductor.
It is indeed widely accepted that, in an isolating (polymer) matrix electrical
charges follow the paths of a conducting filler network. Upon increasing the
concentration of conductive particles (i.e., CB) in a nonconducting matrix an
isolator–conductor transition is thus, expected to occur. In modern versions
of the percolation theory two models are considered, both leading to (expect-
edly) universal scaling relationships for the conductivity and permittivity
with respect to a fractal network of conductive particles.92 The so-called lattice
gas model describes the conductivity as an anomalous diffusion of charge car-
riers on the conducting paths The so-called equivalent circuit model consider the
fractal particles system as a random mixture of tiny resistors and capacitors
assumed to form in the local interphase regions between conductive particles
and the surrounding isolating (polymer) matrix. Note that, in order to sig-
nificantly modify the conductivity of an insulating material it is not required
Polymers and Carbon Black 173

Table 5.11
Typical Coloring Carbon Blacks
Carbon Mean Particle N2 1 Specific CTAB2 DBPA2 Tinting4
Black Type Diameter (nm) Area (m2/g) (m2/g) (cm3/100 g) Strength
HCFa 8.9 430 300 95 133
HCF 1.3 575 330 95 133
HCF 16.4 525 270 98 135
MCFb 16.0 190 175 65 130
MCF 20.0 125 131 56 133
MCF 24.2 95 102 55 121
MCF 29.7 65 66 50 97
LCFc 73.5 33 33 63 64
LCF 85.7 25 29 75 48
1 BET, ASTM D 3037; 2 ASTM D 3765; 3 ASTM D 2414; 4 ASTM D3265.
a High color furnace; b Medium color furnace; c Low color furnace.

to have actual physical contact between dispersed conductive particles; only


close proximity is required because electrons can cross short layers of insulat-
ing materials through so-called tunnel effects.93 Measurements on CB filled
polymer systems by various authors indicate that gap widths of the order of
2–5 nm between neighboring particles are sufficient for quantum tunneling
of electrons, and that the required minimum gap width is independent of the
filler volume fraction.94–96
Color differences in materials depend on the interaction of light with
electrons. In the graphitic layers of CB particles, electrons are free to
vibrate at practically any frequency and therefore absorb all wavelengths
of visible light ranging from infrared to ultraviolet. Table 5.11 gives
the characterization data for a few typical CB grades used in coloring
applications.
As can be seen, there are correlations between the size of elementary par-
ticles, the structure of the aggregate and the coloring capacity of CB, but
the more complex the aggregates the more difficult the ­dispersion process.
The UV absorption capabilities of CB are used to protect polyolefins against
light exposure; a few percentages offer an adequate protection without much
modification of rheological and mechanical properties.

5.2.2 Effect of Carbon Black on Rheological Properties


of Thermoplastics
Essentially the same (qualitative) effects as with rubber materials are
observed when adding CB particles to thermoplastics. As illustrated by
Figure 5.51, the disappearance of the Newtonian plateau and a (apparent)
yield stress limit is observed when adding increasing quantities of CB to
thermoplastic melts.97,98
174 Filled Polymers

LDPE / N220 carbon black 105 Polystyrene / N110 carbon black


3 0%
10 0%
1%
5% 104 5%

Shear viscosity, kPa.s


10% 10%
Shear viscosity, kPa.s

102 20 % 20%
103 25%
101

102
100
Temp.= 150 °C Temp. = 170 °C

101
10–1
100 101 102 103 10–2 10–1 100 101 102
Shear stress, kPa Shear rate, s –1

Figure 5.51
Effect of carbon black loading in the shear viscosity function of LDPE and PS melts. (Drawn
using data by C.Y. Ma, J.L. White, F.C. Weissert, K. Min, J. Non-Newtonian Fl. Mech., 17, 275–287,
1985 and V.M. Lobe, J.L. White, Polym. Eng. Sci., 19, 617–624, 1979.)

105 Polystyrene / N110 carbon black


Extensional viscosity, kPa.s

25%
104
20%

0%
103

Temperature = 170 °C
Strain rate : 0.063 s–1

102
0.1 0.2 0.5 1 2 5 10 20 50 100
Time, s
Figure 5.52
Effect of carbon black loading in the shear viscosity function of LDPE and PS melts. (Drawn
using data by M. Lobe, J.L. White, Polym. Eng. Sci., 19, 617–624.)

Extensional flow properties are important in a number of processing tech-


niques for thermoplastics, such as extrusion blowing and blow molding and
accordingly the effect of CB content of the extensional viscosity has been
somewhat studied by several authors. As shown in Figure 5.52, it is gener-
ally observed that ηE increases with higher CB level, but typical high strain
effects, such as strain hardening, seem to be suppressed or at least moved
outside the experimental window.
Polymers and Carbon Black 175

In fact the effects of CB type and level on the flow properties of


t­ hermoplastics, whilst less documented, appear very similar (at least quali-
tatively) to what is observed with rubber materials. Accordingly the loading
dependence of viscosity can be treated with similar equations, but the com-
plications due to strong polymer–CB interactions seem to be either absent or
at least negligible (to the author’s knowledge).

5.2.3 Effect of Carbon Black on Electrical Conductivity


of Thermoplastics
With conductivities in the 10−5 − 10−19 (Ω.cm)−1 range, polymers are poor elec-
trical conductor, and, as might be expected, a sufficient quantity of CB par-
ticles, whose conductivity is in the 102−105 (Ω.cm)−1 range, is likely to impart
some electrical conductivity properties to composites. It is fairly obvious that
the principal cause of differences in the conductivity of CB–­thermoplastic
composites is the structure formed by CB particles, not the intrinsic conduc-
tivity of the carbon itself. It follows that the structure of the aggregate and
its overall size are important factors and therefore some CB grades are more
suitable than others. Because a network of connected or at least sufficiently
close particles within the thermoplastic matrix is the ­necessary condition
for the composite to exhibit electrical conductivity, the aggregates must be
as “open” as possible with the largest number of elementary particles per
aggregate a favoring factor. Table 5.12 compares several grades of so-called
“conductive” carbon blacks, using characteristics, which may be ­considered
as important with respect to electrical properties of composites.
Specific area is surely an important factor in obtaining a CB network above
a certain level (the so-called percolation level), but acetylene black, with its
large number of particles per aggregate, has a very favorable structure for
imparting a good conductivity to composites, as discussed by Medalia.99
Special, open-branches aggregates grade are produced by CB manufacturers
for preparing conductive filled polymer materials.

Table 5.12
Properties of “Conductive” Carbon Blacks
Elementary Number of
N2 Specific DPBA Particle Size Particles/
Grade Area (m2/g) (cm3/100 g) (nm) Aggregate Anisometry
N472 238 193 22.1 481 2.00
N880 10 31 192.0 4.9 1.21
Acetylene 51 262 42.2 1070 2.00
Vulcan® XC-72
(Cabot) 220 170 30 n.a. n.a.
HiBlack® 420B
(Evonik) 88 n.a. 24 n.a. n.a.
176 Filled Polymers

When the suitable quantity of CB particles is dispersed in a single poly-


mer, the resulting black network is sensitive to processing because of con-
verging flows which may, in the best cases induce strong anisotropy effects
(and likely a good electrical conductivity along the processing flow) or in the
worst cases result in a disruption of the conductive network. This explains
the recent interest for conductive materials prepared by ­loading polymer
alloys with CB. Indeed most incompatible polymer blends are made with
finely dispersed components and therefore, have a co-­continuous heteroge-
neous structure. Because CB has not the same “affinity” for all polymers, a
careful selection of blend components with the appropriate blending proce-
dure allow to somewhat control the localization of particles in the interphase
region,100–103 as illustrated in Figure 5.53.
Preparing conductive polymer blends with selected localization of CB has
attracted considerable research interested with a near continuous stream of
publications whose detailed review is outside of the scope of the present
book. The thermodynamic and kinetic factors that govern the localization of
CB particles at interfaces in polymer blends are rather complex since systems
far from equilibrium state are obtained, as recently considered by several
authors.104,105
Different results are obtained when using short carbon fibers For instance,
Zhanga et al.106 prepared high-density polyethylene (HDPE)/isotactic

1012
1 : Carbon black in PS phase

2 : Carbon black in PE phase


1010
3 : Carbon black in PE phase of
co-continuous PS/PE blend
108
Resistivity (Ohm.cm)

4 3 2 1 4 : Carbon black at interface of


co-continuous PS/PE blend

106
Lowest percolation level

104 Carbon black particles


localize at interfaces
PS
102

PE
100
0 2 4 6 8 10 12 14 16
Carbon black volume fraction, %
Percolation level for system 1
Figure 5.53
Controlling the electrical conductivity through carbon black localization in thermoplastic
blends. (Drawn using data from F. Gubbels, E. Vanlathem, R. Jerome, R. Deltour, Ph. Teyssié.
2nd International Conference on Carbon Black, Mulhouse, France, Sept. 27–30, 1993, 397.)
Polymers and Carbon Black 177

polypropylene (iPP) blends filled with vapor-grown carbon fibers (VGCF).


They observed that the VGCF percolation threshold was only 1.25 phr, a much
lower level than with the individual polymers. SEM micrographs revealed
that the effect is essentially due to the selective location of the carbon fibers in
the HDPE phase. It follows that a double percolation is the basic requirement
for the conductivity of such composites, i.e., first a percolation of carbon fibers
in the HDPE phase, then the continuity of this phase in the polymer blends.
Preparing a blends with incompatible polymers and selectively locating
CB in the interphase region is obviously quite difficult and how the mor-
phology of such complex materials evolves during the processing/shap-
ing operations is generally not well documented. Therefore, it is worth
underlining that, much simpler systems are exhibiting quite interesting
electrical conductive properties that are also assigned to a (natural) selec-
tive location of CB. For instance CB filled poly(ethylene-co-butyl acrylate)
is used in semi-conducting power cable shielding but filler loadings must
be above the critical value of the percolation theory, i.e., Φc > 0.14–0.17 (up to
40% by weight) to ensure the adequate electrical properties and an optimal
dispersion is required. Poly(ethylene-co-alkyl acrylate)s are semicrystalline
materials whose degree of crystallinity depends on the level of acrylate.107
At room temperature; such materials naturally develop an heterogeneous
structure such that crystallites form only in ethylene rich regions. Butyl
acrylate (BA) rich regions are essentially amorphous. In composites made
by mixing CB particles with poly(ethylene-co-butyl acrylate) (EBA), semi-
conducting properties depend not only on the network microstructure of
connected CB particles but also on specific interactions between the filler
and the amorphous BA-rich regions of the matrix. When EBA/CB compos-
ites are hold above their melting temperature for a sufficient time, the filler
microstructure can somewhat evolve, with consequently modified electri-
cal properties after cooling.108 Such systems exhibit also a set of interest-
ing nonlinear viscoelastic properties, not only in the solid (i.e., at room
temperature), but also in the molten state.109 Through a series of advanced
rheological measurements, Leblanc and Jäger110 came to the conclusion that
CB particles essentially concentrate in amorphous regions, leading thus to
a highly complex morphology with at least three phases: PE rich crystal-
lites (no black in it), BA rich amorphous and CB (Figure 5.54). In terms of
modulus, the ranking is obviously CB > crystallites > amorphous. It follows
that, in the solid state, EBA CB composites exhibit viscoelastic properties
that are essentially dominated by the CB rich BA phase, whilst in the mol-
ten state, the PE rich phase contributes significantly to the initial response
in such a manner that a certain strain limit has to be reached before the
filler reinforcing effect in the BA phase is observed. Because of specific
interactions between the acrylate rich regions of the material and carbon
black particles, a complex structure naturally occurs, similar to what can
be obtained by carefully controlling the location of CB in incompatible
polymer blends.
178

RPA-FT; 200°C; 1 Hz; 2 tests (Фblack = 0.275) Behavior in solid and molten states
(Фblack = 0.248)
Solid state PE cristallites (underformable)
100 (Фblack = 0.226)
(Фblack = 0.203)
Virgin EBA
Strain

10
Carbon black/butyl acrylate
Molten state rich amorphous regions

Complex modulus G*, MPa


118.7 Strain
1

10 100
Strain amplitude, % Molten PE rich regions (deformable)
Figure 5.54
Strain sweep tests (rubber process analyzer, updated for fourier transform rheometry) on molten virgin and carbon black filled poly(ethylene-co-butyl
acrylate) composites and likely morphology in both the solid and molten states.
Filled Polymers
Polymers and Carbon Black 179

References
1. A.C. Patel, J.T. Byers. The influence of tread grade carbon blacks at optimum
loadings on rubber compound properties. Rubb. India, 34 (4), 9–13, 1982.
2. S. Montes, J.L. White, N. Nakajima. Rheological behaviour of rubber carbon
black compounds in various shear histories. J. Non-Newtonian Fluid Mech., 28,
183–212, 1988.
3. C. Barrès, J.L. Leblanc. Recent developments in shear rheometry of uncured rub-
ber compounds. part 1: design, construction and validation of a sliding ­cylinder
rheometer. Polymer Testing, 19, 177–191, 2000.
4. W.H. Herschel, R. Bulkley. Konsistenzmessungen von Gummi-Benzollösungen.
Kolloid Z., 39, 291–300, 1926; Measurement of consistency as applied to rubber-
benzene solution. Proc. Am. Soc. Testing Mat., 26, 621–633, 1926.
5. N. Casson. In Rheology of Disperse Systems, C.C. Mill, Ed. Pergamon, London,
UK, 84, 1959.
6. J.L. White, Y. Wang, A.I. Ysayev, N. Nakajima, F.C. Weissert, K. Min.Modeling
of shear viscosity behavior and extrusion through dies for rubber compounds.
Rubb. Chem. Technol., 60, 337–360, 1987.
7. See for instance C.W. Macosko. Rheology, Principles, Measurements and Applications.
VCH Publishers, New York, NY, 416, 1994.
8. A.K. Bagchi, K.K. Sirkar. Extrusion die swell of carbon black filled natural rub-
ber. J. Appl. Polym. Sci., 23, 1653–1670, 1979.
9. G. Kraus. In Science and Technology of Rubber, F.R. Eirich Ed. Acad. Press, London,
UK, 1978.
10. C.G. Robertson, C.M. Roland, J.E. Puskas. Nonlinear rheology of hyperbranched
polyisobutylene. J. Rheol., 46, 307–320, 2002.
11. J.L. Leblanc. Rubber-filler interactions and rheological properties in filled com-
pounds. Prog. Polym. Sci., 27, 627–687, 2002.
12. E.M. Dannenberg. Bound rubber and carbon black reinforcement. Rubb. Chem.
Technol., 59, 512–524, 1986.
13. S. Wolff, M-J. Wang, E-H. Tan. Filler-elastomer interaction. Part VII. Study on
bound rubber. Rubb. Chem. Technol., 66, 163–177, 1993.
14. J. Kruse. Rubber microscopy. Rubb. Chem. Technol., 46, 653–785, 1973. See Figures
29 and 30 in the paper.
15. L.L. Ban, W.M. Hess, L.A. Papazian. New studies of carbon rubber gel. Rubb.
Chem. Technol., 47, 858–894, 1974. See Figure 3 in the paper.
16. S. Kaufman, W.P. Slichter, D.D. Davies. Nuclear magnetic resonance study of
rubber–carbon black interaction. J. Polym. Sci., A2 (9), 829–839, 1971.
17. J. O’Brien, E. Cashell, G.E. Wardell, V.J. Mc Brierty. An NMR investigation of
the interaction between carbon black and cis-polybutadiene. Macromolecules, 9,
653–659, 1976.
18. J.C. Kenny, V.J. McBrierty, Z. Rigbi, D.C. Douglass. Carbon black filled natural
rubber. 1. Structural investigation. Macromolecules, 24, 436–443, 1991.
19. F. Yatsuyanagi, H. Kaidou, M. Ito. Relationship between viscoelastic properties
and characteristics of filler-gel in filled rubber system. Rubb. Chem. Technol., 72,
657–672, 1999.
20. J.L. Leblanc. Insight into elastomer-filler interactions and their role in the ­processing
behaviour of rubber compounds. Plast. Rubb. Proc. Technol., 10 (2), 110–129, 1994.
180 Filled Polymers

21. K. Nakashima, H. Fukuta, M. Mineki. Anisotropic shrinkage of injection—


molded rubber. J. Appl. Polym. Sci., 17, 769–778, 1973.
22. J.L. Leblanc, B. Stragliati. An extraction kinetics method to study the morphology
of carbon black filled rubber compounds. J. Appl. Polym. Sci., 63, 959–970, 1997.
23. A.M. Gessler, W.M. Hess, A.I. Medalia. Reinforcement of elastomers with car-
bon black. Part IV. Interaction between carbon black and polymer. Plast. Rubb.
Proc., 3, 141–156, 1978.
24. G.R. Cotten. Mixing of carbon black with rubber: IV. Effect of carbon black char-
acteristics. Plast. Rubb. Proc. Appl., 7 (3), 173–178, 1987.
25. J.L. Leblanc, P. Hardy. Evolution of bound rubber during the storage of uncured
compounds. Kautsch. Gummi, Kunstst., 44, 1119–1124, 1991.
26. J.L. Leblanc. A molecular explanation for the origin of bound rubber in carbon
black filled rubber compounds. J. Appl. Polym. Sci., 66, 2257–2268, 1997.
27. J.P. Cohen-Addad. Silica-siloxane mixtures. structure of the adsorbed layer: chain
length dependence. Polymer, 30, 1820–1823, 1989; ibid. Sol or gel-like behaviour of
ideal silica-siloxane mixtures : percolation approach. Polymer, 33, 2762–2767, 1992.
28. A. Einstein. Eine neue Bestimmung der Molekuldimensionen. Ann. Physik, 19,
289–306, 1906; ibid. 34, 591–592, 1911); note that the second publication is a cor-
rection by Einstein of a calculation mistake; the theoretical reasoning is made in
the 1906 paper but the right equation to consider is given in the 1911s one.
29. G.B. Jeffery. The motion of ellipsoidal particles immersed in a viscous fluid.
Proc. Roy. Soc. London, A102, 161–179, 1923.
30. W. Kuhn. Über quantitative Deutung der Viskosität und Strömungs doppel-
brechung von Suspensionen. Kolloid. Z., 62, 260–285, 1933.
31. E. Guth. Untersuchungen über die Viskosität von Suspensionen und Lösungen.
I. Über die Viskosität von Suspensionen. Kolloid Z., 74, 147–172, 1936; V. Über den
Einfluss der Brownschen Bewegung auf die Viskosität von Ellipsoidsuspensionen.
Kolloid Z., 75, 15–20, 1936.
32. E. Guth, R. Simha. Untersuchungen über die Viskosität von Suspensionen und
Lösungen. III. Über die Viskosität von Kugelsuspensionen. Kolloid Z., 74, 266–
275, 1936.
33. R. Simha. The influence of Brownian movement on the viscosity of suspension.
J. Phys. Chem., 44, 25–34, 1940.
34. M.L. Huggins. Thermodyamic properties of solutions of long-chain compounds.
Ann. N.Y. Acad. Sci., 43, 1–32, 1942.
35. R. Eisenschitz. Der Einfluss der Brownschen Bewegung auf die Viskosität von
Suspensionen. Z. Physk. Chem., A163, 133–141, 1953.
36. V. Vand. Theory of viscosity of concentrated suspensions. Nature, 155, 364–365, 1945.
37. J.L. White, J.W. Crowder. The influence of carbon black on the extrusion char-
acteristics and rheological properties of elastomers: BR and SBR. J. Appl. Polym.
Sci., 18, 1013–1038, 1974.
38. L.E. Nielsen. Mechanical Properties of Polymers, Figure 7.11 in the paper as the
source of data. Reinhold Publ., New York, NY, 163, 1962; note that in order
to cover a sufficiently large temperature scale, Nielsen data were somewhat
extrapolated towards higher temperature.
39. A.R. Payne, R.E. Whittaker. Low strain dynamic properties of filled rubbers.
Rubb. Chem. Technol., 44, 440–478, 1971.
40. L. Mullins. Effect of stretching on the properties of rubber. Rubb. Chem. Technol.,
21, 281–300, 1948.
Polymers and Carbon Black 181

41. A.I. Medalia. Filler aggregates and their effects on dynamic properties of rub-
ber vulcanizates. Colloques Internationaux du Centre National de la Recherche
Scientifique, CNRS, Le Bischenberg, Obernai, France, Sept. 24–26, 1973. Paper
#231, 62–79, CNRS, Paris, 1975. ISBN 2-222-01749-1.
42. A.I. Medalia. Morphology of aggregates. IV. Effective volume of aggregates of
carbon black from electron microscopy; application to vehicle adsorption and to
die swell of filled rubber. J. Colloid Interface Sci., 32, 115–131, 1970. Equation 31.
43. M.J. Wang, S. Wolff, E.H. Tan. Filler-elastomer interactions. Part VIII. The role of
the distance between filler aggregates in the dynamic properties of filled vulca-
nizates. Rubb. Chem. Technol. 66, 178–195, 1993.
44. I. Soos. Rheological characterization of carbon black filled elastomers. Kautsch.
Gummi, Kunstst., 47, 502–507, 1994.
45. L. Mullins, N.R. Tobin. Stress softening in rubber vulcanizates. Part I. Use of s
strain amplitude factor to describe the elastic behaviour of filler-reinforced vul-
canized rubber. J. Appl. Polym. Sci., 9, 2993–3009, 1965.
46. H. Nukaga, S. Fujinami, H. Watabe, K.Nakajima, T. Nishi. Nanorheological
analysis of polymer surfaces by atomic force microscopy. Jpn J. Appl. Phys., 44,
5425–5429, 2005.
47. A.F. Blanchard. Carbon particle reinforcement in rubber. Rubb. J., 153 (2), 44–53,
1971; ibid, 153 (3/4), 25–38, 66, 1971.
48. J.M. Caruthers, R.E. Cohen, A.I. Medalia. Effect of carbon black on hysteresis
of rubber vulcanizates : equivalence of surface area and loading. Rubb. Chem.
Technol., 49, 1076–1094, 1976.
49. A.I. Medalia, S.G. Laube. Influence of carbon black surface properties and morphol-
ogy on hysteresis of rubber vulcanizates. Rubb. Chem. Technol., 51, 89–109, 1978.
50. G.B. Ouyang, N. Tokita, M.J. Wang. Hysteresis mechanisms for carbon black
filled vulcanizates – a network junction theory for carbon black reinforcement.
ACS, Rubb. Div. Mtg, Cleveland, OH, Oct. 17–20, 1995. Paper 108.
51. G.B. Ouyang. Modulus, hysteresis and the Payne effect. Network junction model
for carbon black reinforcement. Kautch. Gummi Kunstst., 59 (6), 332–343, 2006.
52. A.N. Gent, B. Park. Compression of rubber layers bonded between two parallel
rigid cylinders or between two rigig spheres. Rubb. Chem. Technol., 59, 77–83, 1986.
53. A.N. Gent, Y.C. Hwang. Elastic behavior of a rubber layer bonded between two
rigid spheres. Rubb. Chem. Technol., 61, 630–638, 1988.
54. M.J. Wang. Effect of polymer–filler and filler–filler interactions on dynamic
properties of filled vulcanizates. Rubb. Chem. Technol., 71, 520–589, 1998.
55. J.T. Byers, A.C. Patel. Carbon black’s influence on tread wear and hysteresis.
Rubb. World., 188 (3), 21–32, 1983.
56. F.C. Welsch, B.R. Richmond, C.B. Reach, R.J. Emerson. Dynamic reactive mixing
of carbon black in tire treads. ACS, Rubb. Div. Mtg., Philadelphia, PA, May 2–5,
1995. Paper 59.
57. T. Yamaguchi, I. Kurimoto, K. Ohashi, T. Okita. Novel carbon black/rubber
­coupling agent. Kautsch. Gummi Kunstst., 42, 403–409, 1989.
58. See for instance G. Kraus, C.W. Childers, K.W. Rollmann. Stress softening in car-
bon black-reinforced vulcanizates. Strain rate and temperature effects. J. Appl.
Polym. Sci., 10, 229–244, 1966; G. Kraus. Mechanical losses in carbon black filled
rubbers. J. Appl. Polym. Sci.; Appl. Polym. Symp., 39, 75–92, 1984.
59. G. Heinrich, T.A. Vilgis. Contribution of entanglements in the mechanical proper-
ties of carbon black filled polymer networks. Macromolecules, 26, 1109–1119, 1995.
182 Filled Polymers

60. P.S. Theocaris. Concept of Mesophase in Composites. Springer series, Polymers–


Properties and Applications. Springer Verlag, New York, NY, 1987.
61. P.S. Theocaris, G.D. Spathis. Glass-transition behavior of particle composites
modeled on the concept of interphase. J. Appl. Polym. Sci., 27 (8), 3019–3025, 1982.
62. A.R. Payne. The dynamic properties of carbon black loaded natural rubber vul-
canizates. Part I. J. Appl. Polym. Sci., 6, 57–63, 1962; Part II. J. Appl. Polym. Sci., 6,
368–372, 1962.
63. A.R. Payne. Strainwork dependence of filler-loaded vulcanizates. J. Appl. Polym.
Sci., 8, 2661–2686, 1963.
64. A. R. Payne. Reinforcement of Elastomers. G. Kraus, Ed. Wiley Interscience, New
York, NY, 1965.
65. W.P. Fletcher, A.N. Gent. Non-linearity in the dynamic properties of vulcanised
rubber compounds. Trans. Inst. Rubb. Ind. 29, 266–280, 1953.
66. C.M. Roland. Dynamic mechanical behavior of filled rubber at small strain.
J. Rheol., 34, 25–34, 1989.
67. M.J. Wang, W.J. Patterson, G.B. Ouyang. ACS Rubb. Div. Mtg, Montreal, Canada,
May 5–8, 1996. Paper #33; abstract in Rubb. Chem. Technol., 69, 722, 1996; Dynamic
stress-softening of filled vulcanizates. Kautch. Gummi Kunstst., 51, 106–117, 1998.
68. J.A.C. Harwood, A.R. Payne. Stress softening in natural rubber vulcanizates.
Part IV. Unfilled vulcanizates. J. Appl. Polym. Sci., 10 (8), 1203–1211, 1966.
69. A.R. Payne. The role of hysteresis in polymers. Rubb. J., 146, 36–49, 1964.
70. G. Kraus. Mechanical losses in carbon black filled rubbers. J. Appl. Polym. Sci.,
Appl. Polym. Symp. 39, 75–92, 1984.
71. B.B. Mendelbrot. Les objets fractals—Forme, hasard et dimensions, 4th Ed. Champs,
Flammarion, Paris, 1995. ISBN 2-08-081301-3; translation in English: Fractals,
Form, Chance and Dimensions. W.H. Freeman & Co, Springer, The Netherlands.
ISBN 0716704730.
72. G. Huber, T. A.Vilgis, G. Heinrich. Universal properties in the dynamical defor-
mation of filled rubbers. J. Phys.: Condens. Matter, 8 (29), L409, 1996.
73. G. Kraus. Mechanical losses in carbon-black filled rubbers. J. Appl. Polym. Sci.,
Appl. Polym. Symp., 39, 75–92, 1984.
74. J.D. Ulmer. Strain dependence of dynamic mechanical properties of carbon
black filled compounds. Rubb. Chem. Technol., 69, 15–47, 1996.
75. M. Gerspacher, C.P. O’Farrell, C. Tricot, L. Nikiel, H.A. Yanga. Modeling of vis-
coelastic properties of filled rubber compounds. ACS Rubb. Div; Mtg, Louisville,
KY, Oct. 8–11, 1996. Paper 74.
76. F. Clément, L. Bokobza, L. Monnerie. Investigation of the Payne effect and its
temperature dependence on silica-filled polydimethylsiloxane networks. Part
II: Test of quantitative models. Rubb. Chem. Technol., 78, 232–244, 2005.
77. M. Klüppel, G. Heinrich. Fractal structures in carbon black reinforced rubbers.
Rubb. Chem. Technol., 68, 623–651, 1995.
78. M. Klüppel, R. Schuster, G. Heinrich. Structure and properties of reinforcing
fractal filler networks in elastomers. Rubb. Chem. Technol., 70, 243–255, 1997.
79. G. Heinrich, M. Klüppel. Recent advances in the theory of filler networking in
elastomers. Adv. Polym. Sci., 160, 1–44, 2002.
80. M. Klüppel. The role of disorder in filler reinforcement of elastomers on various
length scales. Adv. Polym. Sci., 164, 1–86, 2003.
81. A. van de Walle, C. Tricot, M. Gerspacher. Modeling carbon black reinforcement in
rubber compound. ACS, Rubb. Div. Mtg., Pittsburgh, PA, Oct. 11–14, 1994. Paper 10.
Polymers and Carbon Black 183

82. M. Gerspacher, C.P. O’Farrell, H.H. Yang, C. Tricot. Modeling of the carbon black
reinforcement mechanism in elastomers. Rubb. World, 214 (3), 27, 30, 49, 1966.
83. A. Lion. Strain-dependent dynamic properties of filled rubber: a non-linear
viscoelastic approach based on structural variables. Rubb. Chem. Technol., 72,
410–429, 1999.
84. A. Lion, C. Kardelky, P. Haupt. On the frequency and amplitude dependence of
the Payne effect : theory and experiments. Rubb. Chem. Technol., 76, 533–547, 2003.
85. J.M. Funt. Dynamic testing and reinforcement of rubber. Rubb. Chem. Technol.,
61, 842–865, 1988.
86. H. Montes, F. Lequeux, J. Berriot. Influence of the glass transition tempera-
ture gradient on the nonlinear viscoelastic behavior in reinforced elastomers.
Macromolecules, 36, 8107–8118, 2003.
87. I. Langmuir. The adsorption of gases on plane surfaces of glass, mica and plati-
num. J. Am. Chem. Soc., 40, 1361–1403, 1918.
88. P.G. Maier, D. Göritz. Molecular interpretation of the Payne effect. Kautsch.
Gummi, Kunstst., 49, 18–21, 1996.
89. L. Ladouce-Stelandre, Y. Bomal, L. Flandrin, D. Labarre. Dynamic mechanical
properties of precipitated silica filled rubber: influence of morphology and cou-
pling agent. Rubb. Chem. Technol., 76, 145–159, 2005.
90. M.J. Wang. Effect of polymer-filler and filler-filler interactions on dynamic prop-
erties of filled vulcanizates. Rubb. Chem. Technol., 71, 520–589, 1998.
91. M.I. Aranguren, E. Mora, C. W. Macosko, and J. Saam. Rheological and mechanical
properties of filled rubber : silica-silicone. Rubb. Chem. Technol., 67, 820–833, 1994.
92. Y. Gefen, A. Aharony, S. Alexander. Anomalous diffusion on dercolating clus-
ters. Phys. Rev. Lett., 50, 77–80, 1983.
93. E.K. Sichel, J.I. Gittleman. Tunneling conduction in carbon-polymer compos-
ites. In Carbon Black–Polymer Composites, the Physics of Electrically Conducting
Composites, E.K. Sichel Ed. Marcel Dekker, New York, NY, 51–77, 1982.
94. P. Sheng, E. K. Sichel, J. I. Gittleman. Fluctuation-induced tunneling conduction
in Carbon-Polyvinylchloride composites. Phys. Rev. Lett., 40, 1197–1200, 1978.
95. J.G. Meier, M. Klüppel. Carbon black networking in elastomers monitored by
dynamic mechanical and dielectric spectroscopy. Macromol. Mater. Eng., 293,
12–38, 2008.
96. J.G. Meier, J.W. Mani, M. Klüppel. Analysis of carbon black networking in elas-
tomers by dielectric spectroscopy. Phys. Rev. B, 75, 054202, 2007.
97. C.Y. Ma, J.L. White, F.C. Weissert, K. Min. Flow patterns in carbon black filled poly-
ethylene at the entrance to a die. J. Non-Newtonian Fl. Mech., 17, 275–287, 1985.
98. V.M. Lobe, J.L. White. An experimental study of the influence of carbon black on
the rheological properties of a polystyrene melt. Polym. Eng. Sci., 19, 617–624, 1979.
99. A.I. Medalia. Nature of carbon black and its morphology in composites. In
Carbon Black–Polymer Composites, the Physics of Electrically Conducting Composites,
E.K. Sichel Ed. Marcel Dekker, New York, NY, 1–49, 1982.
100. F. Gubbels, E. Vanlathem, R. Jerome, R. Deltour, Ph. Teyssié. Selective control of
the localization of carbon black in polymer blends. 2nd International Conference on
Carbon Black, Mulhouse, France, Sept.27–30, 1993, 397.
101. F. Gubbels, R. Jerome, Ph. Teyssié, E. Vanlathem, R. Deltour, A. Calderone, V.
Parente, J.L. Bredas. Selective localization of carbon black in immiscible polymer
blends: A useful tool to design electrical conductive composites. Macromolecules,
27, 1972–1974, 1994.
184 Filled Polymers

102. F. Gubbels, S. Blacher, E. Vanlathem, R. Jerome, F. Brouers, Ph. Teyssié. Design of


electrical conductive composites : key role of the morphology on the electrical prop-
erties of carbon black filled polymer blends. Macromolecules, 28, 1559–1566, 1995.
103. C. Calberg, S. Blacher, F. Gubbels, F. Brouers, R. Deltour, R. Jerome. Electrical and
dielectrical properties of carbon black filled co-continuous two phase polymer
blends. J. Phys. D. Appl. Phys., 32, 1517–1525, 1999.
104. A.E. Zaikin, E.A. Zharinova, R.S. Bikmullin. Specifics of localization of carbon
black at the interface between polymer phases. Polym. Sci. Series A, 49 (3), 328–
336, 2007.
105. Zhongbin Xu, Chao Zhao, Aijuan Gu, Zhengping Fang , Lifang Tong. Effect of
morphology on the electric conductivity of binary polymer blends filled with
carbon black. J. Appl. Polym. Sci., 106 (3), 2008–2017, 2007.
106. C. Zhanga, X.S. Yib, H. Yuic, S. Asaia, M. Sumitaa. Selective location and double
percolation of short carbon fiber filled polymer blends: high-density polyethyl-
ene/isotactic polypropylene. Mat. Lett., 36 (1–4), 186–190, 1998.
107. J.F. Feller, I. Linossier, S. Pimbert, G. Levesque. Carbon black filled poly(ethylene-
co-alkyl acrylate) composites : calorimetric studies. J. Appl. Polym. Sci., 79, 779–
793, 2001.
108. K.-M. Jäger, D.H. McQueen. Thermal stabilities of electrical properties of EBA/
CB composites. Kautsch. Gummi, Kunstst., 52, 734–741, 1999.
109. K.-M. Jäger, S.S. Eggen. Scaling of the viscoelasticity of highly filled carbon black
polyethylene composites above the melting point. Polymer, 45, 7681–7692, 2004.
110. J.L. Leblanc, K.-M. Jäger. Investigating non-linear viscoelastic properties of molten
carbon black/poly(ethylene-co-butyl acrylate) composites, using Fourier transform
rheometry and other test techniques. J. Appl. Polym. Sci., 101, 4071–4082, 2006.
111 E.M. Dannenberg. The effect of filler surface modification on elastomer rein-
forcement. Colloques Internationaux du Centre National de la Recherche
Scientifique, CNRS, Le Bischenberg, Obernai, France, Sept. 24–26, 1973, Colloque
# 231, 129–135, CNRS, Paris, 1975. ISBN 2-222-01749-1.
112. G. Cotten. Influence of carbon black on processability of rubber stocks. I. Bound
rubber formation. Rubb. Chem. Technol., 48, 548–557, 1975.
113. G. Cotten. Mixing of carbon black with rubber. II. Mechanism of carbon black
incorporation. Rubb. Chem. Technol., 58, 774–784, 1985.
114. J.J. Brennan, T.E. Jermyn, B.B. Boonstra. Carbon black-polymer interaction: a
measure of reinforcement. J. Appl. Polym. Sci., 8, 2687–2706, 1964.
115. G.R. Cotten, E.M. Dannenberg. A method for evaluation of carbon blacks and
correlation with road wear ratings. Tire Sci. Technol., 2 (3), 211–227, 1974.
116. S. Montes, J.L. White. A comparative rheological investigation of natural and
synthetic cis-1,4 polyisoprenes and their carbon black compounds. Rubb. Chem.
Technol., 55, 1354-1369, 1982.
117. H.J. Song, J.L. White, K. Min, N. Nakajima, F.C. Weissert. Rheological proper-
ties, extrudate swell, and die extrusion flow marker experiments for rubber car-
bon black compounds. Adv. Polym. Technol., 8, 421-449, 1988.
118. K.C. Shin, J.L. White, R. Brzoskowski, N. Nakajima. Rheological behavior and
extrusion shrinkage of rubber-carbon black compounds. Kautch. Gummi Kunstst.,
43, 181-188 (1990).
119. K.C. Shin, J.L. White, N. Nakajima. Extrudate character and post-extrusion
shrinkage of rheologically characterized rubber-carbon black compounds and
their interpretation. J. non-Newtonian Fluid Mech., 37, 95-108, 1990.
Polymers and Carbon Black 185

Appendix 5

A5.1  Network Junction Theory


[M.J. Wang, S. Wolff, E.H. Tan. Rubb. Chem. Technol., 66, 178–195, 1993. G.B.
Ouyang, N. Tokita, M.J. Wang. ACS, Rubb. Div. Mtg., Cleveland, OH, Oct.
17–20, 1995. Paper 108.]

A5.1.1  Developing the Model


Aggregates consist of Np elementary particles of diameter d, and are assimi-
lated to equivalent spheres of diameter D. At the end point of the crushed
DBP absorption measurement, the volume of the absorbed DBP is the sum
of the void volume within the aggregates and the volume between equiva-
lent spheres. The volume fraction of equivalent spheres in the system differs
according to the type of packing. The loosest packing for a touching sphere
system is the cubic arrangement, for which the theoretical volume fraction
of spheres is π  = 0.524. Compact cubic arrangement of CB aggregates (cDBP
6
absorption test)
d
In a rubber matrix,
D CB aggregates keep the
same arrangement as in
DBP but junction points
hgap
are widened by a dis-
tance hgap.

gm
Data: ρ: = 1.86. : filler density Np = number of elementary ­particles
cm 3 (CB) in an aggregate
d = elementary particle diameter
D = equivalent sphere diameter

103 ⋅ cm 3 cm 3
cDBP : = 0.87 ⋅ : crushedDBP absorption (here N330)   cDBP = 0.87 
kg gm

MCB : = 150 . gm: mass of CB sample


186 Filled Polymers

Maximum filler volume fraction at the end of cDBP absorption test:

MCB Note: such a ­maximum


volume fraction is
ρ
Φ max : = [1] Φ max = 0.382 quite conform to the
MCB
+ cDBP ⋅ MCB cubic arrangement of
ρ spheres.

The overall volume of the cDBP + CB mixture is thus [MCB(ρ−1 + cDBP)]3


If one considers that, in a rubber matrix, CB aggregates keep the sim-
ple cubic arrangement, the distance between neighboring aggregates is
expanded by a distance hgap, such that for a fraction Φ of equivalent spheres
of diameter D, one has:

3 π −1

hgap =  ⋅ Φ 3 − 1 ⋅ D
 6 

The overall volume of a rubber + CB compound is thus:

3
 1 1

  1
 M ⋅ + cDBP   3  ζ  3

+ h ⋅
gap  Na ⋅
  CB  ρ  
  CB 2   [2]
 

where NaCB is the number of aggregates in MCB grams of CB:

ζ MCB ζ
NaCB ⋅ = ⋅ [3]
2 π ⋅ d3 2
Np ⋅ ⋅ρ
6

ζ = average number of contact points between neigh-


boring aggregates (minimum is 2 obviously)
By substituting the denominator of the right member of [1] by [2] one
obtains thus, an expression for the volume fraction of CB in a filled com-
pound, i.e.:

MCB MCB
ρ ρ
φ CB = 3
= 3
 1 1
  1 1

  M ⋅  1 + cDBP  3 + h ⋅  Na ⋅ ζ  3    M ⋅  1 + cDBP  3 + h ⋅  3 ⋅ ζ ⋅ MCB  3 
 CB    gap      
 2    Nρ ⋅ π ⋅ d 3 ⋅ ρ  
CB
ρ   CB
ρ  gap
   
Polymers and Carbon Black 187

1
φ CB = 3
 1 1

  1  3  3.ζ 3
ρ ⋅   + cDBP  + hgap ⋅ 
 ρ   N p ⋅ π ⋅ d 3 ⋅ ρ  
 

1
Using [1] it comes: φ CB = 3
 1 1

  1  3  3 .ζ  3

 φ max  + hgap ⋅  N p ⋅ π ⋅ d 3  
 

−1
 1 1

 3.ζ  3
  1 3  1 3
 = > Aggregate junction gap width: hgap = d ⋅ ⋅  −   [4]
 N ⋅ π 
p
 φ CB   φ
max 
 

The elementary particle diameter is not an easily accessible parameter and


even tedious TEM analysis gives at best average values. However if one con-
siders that an aggregate consists of a simple assembly of touching elemen-
tary particles, one has the following relationship between the specific surface
area (in m2/g) and the elementary particles diameter d:

6
Ssp =
d ⋅ ρCB

−1
 3 ⋅ζ   1 1

6 3
  1 3  1 3
hgap = ⋅ ⋅   – 
Therefore: ρCB ⋅ Ssp  N ⋅ π 
p
 φ CB   φ max   [5]
 

The NJ model leads to several interesting conclusions, i.e. for a given CB


grade (i.e. specific surface area and number of primary particles):

1. The junction gap with decreases with increasing number of contact


points
2. The junction gap increases with the number of particles per
aggregate
188 Filled Polymers

A5.1.2  Typical Calculations with the Network Junction Model

i := 0 ⋅⋅6 DBPA cDBP CTAB d


dm 3 /kg dm 3 /kg m 2 /g n−m
 N110   1.14   0.98  125.45   18.32 
         
 N220   1.15   0.98  111.56   20.95 
 N330   1.02   0.87   80.54   30.50 
    cm 3   cm 3   m2  
 N550  DBPA : = 1.17  . gm cDBP : =  0.84  . CTAB : =  41.32  ⋅ d : =  50.58  ⋅ 10−9 ⋅ m
 N660   0.94   0.73  gm  37.52  gm  59.28 
         
 N774   0.74   0.63   30.37   87.20 
         
 N990   0.38   0.37   9.70   294.90 

Estimating the number of particles/aggregate: ε: = 0.847

 2

N pi : = round [1.333 ⋅ (1 + DBPA i ⋅ 102 ⋅ ρ ⋅ 0.0115)] 3⋅ε − 2 , 0   278   N110 
   285   N220 
   
 209   N330 
   
Rem: DBPA (dm3/kg) × 100 = DBPA (cm3/100 g) N p =  298   N550 
 170   N660 
   
 97   N774 
 26   N990 
   
Maximum volume fraction in cDBP absorption test:

 0.354   N110 
 0.354   N220 
   
MCB  0.382   N330 
ρ    
Φ maxi : = Φ max =  0.39   N550 
MCB
+ cDBPi ⋅ MCB  0.424   N660 
ρ    
 0.46   N774 
 0.592   N9990 
   
Number of aggregates in MCB grams of CB:

 9.011.1016   N110 
 5.877.1016   N220 
   
 2.597.1016   N330 
MCB    
NaCBi : = ‘ NaCB =  3.994.1015   N550 
π ⋅ (di )3
N pi ⋅ ⋅ρ  4.349.1015   N660 
6    
 2.395.1015   N774 
 2.31.1014   N990 
   
Polymers and Carbon Black 189

Junction gap width:   ΦCB: = 0.20   : filler volume fraction


A hypothesis is necessary about the number of junctions per aggregate,
let’s say ζ: = 6

−1
 27.827   N110 
 3 ⋅ζ  3
6 3 1 3
1    31.552   N2220 
hgapi :=  ⋅ ⋅ −
 π ⋅ N pi  ρ ⋅ CTAB i  Φ CB Φ maxi     
 44.061   N330 
   
hgap =  99.537  10−9 ⋅ m  N550 
 100.878   N660 
   
 113.176   N774 
   
 285.892   N990 

Let us consider a given CB, i.e., N330

   i: = 3    CTAB: = CTAB3     Φmax: = Φmax3     ΦCB:  =  0.20

   ζ: = 2..10   Np: = 10..300   −1
 3 ⋅ζ  3
6
3
 1
3
1 
hg (ζ , N p ): =  ⋅ ⋅ −
 π.N p  ρ ⋅ CTAB  Φ CB Φ max 

N330, 20% vol. fraction N330, 20% vol. fraction


150 150
Junction gap width, mm
Junction gap width, mm

100

100
50

0
0 100 200 300
500 5 10 Number of primary particles/aggregate
Number of contact points
2 junctions
100 particles 5 junctions
200 particles 10 junctions
300 particles
190 Filled Polymers

A5.1.3  Strain Amplification Factor from the Network Junction Theory


A5.1.3.1 Modeling the Elastic Behavior of a Rubber Layer between
Two Rigid Spheres
When a compressive force F provokes a small displacement Δx of one sphere
toward the other, there is a compression stiffness which consists of two
parts:
- The rubber layer compressed between the two spheres
- The restraints at the bonded surfaces of the spheres

∆x

D
One has: F = F1 + F2 F F
x
h D+ h
then, with: A = 1 + =
D D

Gent and Park derive the following equations:

F1  π   A  
= E0 ⋅ D ⋅   ⋅  A ⋅ ln  –1 [Gent and Park, Equation 5]
∆x  2   A –1  

F2  π  1 1  1 
= E0 ⋅ D ⋅   ⋅  3 + + + 3.A ⋅ ln  1 −   [Gent and Park,
∆x  8  2 ⋅ A A –1  A
                           Equation 8]

F  π   A    π  1 1  1 
=   ⋅  A ⋅ ln   − 1 +   ⋅  3 + + + 3 ⋅ A ⋅ ln  1–  
E0 ⋅ D ⋅ ∆x  2    A − 1   8   2 ⋅ A A –1  A

F π   A  5 ⋅ A –2 ⋅ A 2 –1 
which simplifies to: = ⋅  A ⋅ ln  +
E0 ⋅ D ⋅ ∆x 8   A –1  2 ⋅ A ⋅ ( A –1) 

Gent and Hwang write a different equation for the same problem:

F π   A  1 1 
= E0 ⋅ D ⋅ ⋅  A ⋅ ln   –1 + + [Gent and Hwang,
(E0 ⋅ ∆x ⋅ D) 8   A –1  2 ⋅ A A –1  Equation 3]
Polymers and Carbon Black 191

which, after rearrangement, leads to the same equation:


F π   A  5 ⋅ A –2 ⋅ A 2 –1 
= ⋅  A ⋅ ln  +
E0 ⋅ D ⋅ ∆x 8   A –1  2 ⋅ A ⋅ ( A –1)  [6]
h D+h F π  h   (D + h)  2 ⋅ D2 + D ⋅ h – 2 ⋅ h2 
A = 1+ = => = ⋅  1 +  .ln  +
D D E0 ⋅ D ⋅ ∆x 8  D   h  2 ⋅ (D + h) ⋅ h 

A5.1.3.2  Experimental Results vs. Calculated Data

Exp [1] Calc [1] Calc [2] FEM [2] Sources of data
[1] Measurement with two
F/E0 D∆x F/E0 D∆x F/E0 D∆x F/E0 D∆x spheres of D=41.7 mm spaced
with cured silicone rubber
h/D h/D h/D h/D and calculations with model;
 0.064 7.2 0.064 7.1 0.01 40.90 0.01 40.77  A.N. Gent and B. Park, Rubb.
  Chem. Technol., 59, 77, 1986.
 0.069 5.1 0.069 3.63 0.02 21.01 0.02 20.916 
[2] Calculated with model
 0.075 6.02 0.075 6.15 0.05 8.904 0.05 8.704  and using a finite-element
 
 0.14 3.6 0.14 3.52 0.1 4.749 0.1 4.519  method; A.N. Gent and Y.C.
Data-:  0.2219 2.54 0.219 2.38 0.2 2.579 0.2 2.399  Hwang, Rubb. Chem. Technol.,
    61, 630, 1988.
 0.235 1.99 0.235 2.24 0.5 1.171 0.5 1.122 
 
 0.323 1.6 0.323 1.71 1 0.643 1 0.648 
 0.518 0.929 0.518 1.135 2 0.347 2 0.358 
 
 0.674 0.957 0.674 0.905 2 0.347 2 0.358 

Extracting data for nonlinear regression


100
x1: = stack(Data<0> , Data<2> )

x 2: = stack(Data<4> , Data<6> )
10
x: = stack( x1,x 2)
Stiffness

y 1: = stack(Data<1> , Data<3> )
1
y 2: = stack(Data <5> , Data<7> )

y: = stack( y 1,y 2) 0.1


0.01 0.1 1 10
h/D

With log scales, data exhibit an inverse


dependence so that one can consider a –β
F π  h
much simpler approximate model :   = α ⋅ ⋅      [7]
E0 ⋅ D ⋅ ∆x 8  D
Initial guessed parameters for nonlinear fitting  π ( − C)1 
 C0 ⋅ 8 ⋅ x 
1.2   
C : =    Function to fit with partial derivatives:  F( x ,C): =  π - C1 
⋅x
 0.8   8 
 
π
 −C0 ⋅ ⋅ x − C1 ⋅ ln( x)
 8 
192 Filled Polymers

Res: = GenFit (x, y, C, F)   : calling nonlinear fitting algorithm


π Res1
Z: = Re s 0 , x-    < = fitting equation
8
 1.223 
R2: = corr(Z,y)   < = correlation coefficient r2   Res =      R2 = 0.998
 0.964 

100

10
A simple power law with a negative
Stiffness

exponent and the appropriate pre-


1 factor fits well the data in the h/D
range of interest.
0.1
0.01 0.1 1 10
h/D

A5.1.3.3 Comparing the Theoretical Model with the Approximate


Fitted Equation

D: = 100 h: = 0.001, 0.002..10 α: = Res0 β: = Res1

π  h  (D + h)  2 ⋅ D2 + D ⋅ h – 2 ⋅ h2     π  h  –β 
Z1( h,D): = ⋅  1 +  ⋅ ln  + Z 2 ( h , D ) : = α ⋅  ⋅  
8  D  h  2 ⋅ (D + h) ⋅ h 
 8  D  

1.105 1000
Difference (theory-fitting)

1.104

1.103
Stiffness

500
100

10

1 0
1.103 0.01 0.1 1 10 1.10–3 0.01 0.1 1 10
h h
D = 100 (Theory) D = 100 D = 200
D = 100 (Fitting) Except at very low separation gap, the empirical
D = 200 (Theory) fitting equation gives similar results to the theoretical
D = 200 (Fiting) model in the layer thickness extending down to
one-tenth of the sphere diameter.
Polymers and Carbon Black 193

A5.1.3.4  Strain Amplification Factor


Let us consider a volume V of filled compound: the overall number N of CB
aggregates in this sample is given by:

N NaCB
= Φ CB ⋅ ρ ⋅
V MCB
2 2
With respect to Equation [3], in a cross
 N  3  6 ⋅ Φ CB  3 1
section of the volume V, the number of    =  ⋅      [8]
aggregates is:
V  π ⋅ N p  d 2

When the sample is strained in one direction x by an amount ΔX, the junc-
tions across a section perpendicular to the stretching direction contribute to
the overall stress, in addition to the contribution of the rubber matrix. With
respect to Equation [6], it follows that, over the cross section, the junctions
contribute to a force given by:
2 −β
 N3 π  h gap  where Des is the diameter of the
FJ =   ⋅ ( E0 ⋅ Des ⋅ ∆x ) ⋅ α ⋅ ⋅  equivalent sphere of an aggregate
V 8  Des 

Obviously ΔX = Δx (N/V)1/3 and the 1 −β


network junctions contribute to the  FJ  N  3 π  hgap 
= ⋅ (E0 ⋅ Des ) ⋅ α ⋅ ⋅ 
modulus by a quantity: ∆X  V  8  Des 

From Medalia’s flocε simulation (see Chapter 4, section 4.1.4, Equation 4.6),
one has: Des = d. N p 2
with Np the number of particles of diameter d in an aggregate, and
ε = 0.847
Substitutions are made with respect to Equations [4] and [8]:
−1 –β
 1 
 1

 d.  3 ⋅ ζ  3 ⋅  1  3 –  1  3  
1   N p .π   
 Φ CB   
 Φ max   
FJ  6.Φ CB  3 1   ε
 π   
  = ⋅ ⋅ E0 ⋅  d ⋅ N p   ⋅ α ⋅ ⋅ 
3
∆X  π ⋅ N p  d    8 ε 
 d ⋅ Np 2 
 
 
 

1 –β
 6 ⋅ Φ CB  3 1    3 ⋅ζ 
−1/3  1 1

 ε
 π  −2ε   1  3  1  3 
  => ⋅ ⋅ E ⋅ d ⋅ N ⋅ α ⋅ ⋅ Np ⋅ ⋅  –
    Φ CB   Φ   
2
 π ⋅ N  d  0 p
8   N ⋅ π 
p  p max
  

1+β –β
ε⋅(β + 1)
β
 1

FJ π  6 ⋅ Φ CB  3
 ζ  3   Φ CB  3 
=> = E0 ⋅ α ⋅ ⋅  ⋅ Np 2 ⋅   ⋅ 1–  
∆X 8  π ⋅ N p   2    Φ max  
 
194 Filled Polymers

Over the cross section, there also a contribution from the matrix rubber
alone, so that the resulting stiffness can be estimated as:

FR  2

= E0 ⋅  1–Φ CB 3 
∆X  

F F F
The overall stiffness is thus, = J + R
∆X ∆X ∆X

1+β –β
β
 1

π  6 ⋅ Φ CB  3 ε ⋅(β + 1)
 ζ  3   Φ CB  3  2
  => ECpd = E0 ⋅ α ⋅ ⋅  ⋅ Np 2 ⋅   ⋅ 1–   + E0 ⋅ (1 – Φ CB 3 )
8  π ⋅ N p   2    Φ max  
 

So that the strain amplification factor Xf due to CB presence is:

1+β –β
ε⋅(β +1)
β
 1

  ECpd π  6 ⋅ Φ CB  3
 ζ  3   Φ CB  3   2

= Xf = α ⋅ ⋅  ⋅ Np 2 ⋅   ⋅ 1–   +  1 – Φ CB 3  [9]
E0 8  π ⋅ N p   2    Φ max    
 

A5.1.4 Comparing the Network Junction Strain


Amplification Factor with Experimental Data

G′ [A ] E ′ [B ] E ′ [B ] Source of data:
[A] A.R. Payne, R.E. Whitakker.
Rubb. Chem. Technol., 44, 440–478,
Φ CB Φ CB Φ CB
1971. Figure 3; data on butyl rub-
N330 N347 N327 ber cpd, at 0.05% strain.
[B] J.M. Caruthers, R.E. Cohen,
 0 0.2 0 5.625 0 5  A.I. Medalia. Rubb. Chem. Technol.,
  49, 1076–1094, 1976. Figure 6; data
 0.048 0.4 0.086 6.875 0.086 6.25  on SBR 1500 cpds, with N347 and
 0.092 0.7 0.159 10.781 0.159 6.875  N327 carbon blacks, at 60°C, 0.25
  Hz, 5% strain; note that the last
Data2: =  0.132 1.2 0.175 9.531 0.22 8.906  points in columns 4 and 5 are
 0.168 2 0.22 17.1875 0.239 10.625  cubic interpolation data, in order
  to complete the matrix.
 0.202 4.5 0.274 29.375 0.274 14.0625 
 
 0.232 9.8 0.248 22.4 0.191 7.5 

Strain amplication factor according to Guth and Gold:


Φ: = 0,0.01..0.25
XfGG (Φ) : = (1 + 2.5 ⋅ Φ + 14.1 ⋅ Φ 2 )

Strain amplication factor according to NJ theory:


Data used : i: = 0 .. 2
Polymers and Carbon Black 195

gm
ρ = 1.86 ϒ 3 ζ: = 2 DBP cDBP N p Φ max
cm
1.02 0.87 209 0.390   N330  Np and Φmax were calcu-
α = 1.223   β = 0.964   TT: = 1.22 0.97 333 3 0.357 
 
 N347 
 lated as in above using
 0.96 0.68 179 0.442   N3227  the corresponding DBP
and cDBP data

  Np: = TT<2>   Φmax: = TT<3>

–β
 1+β
ε .(β +1)
β
 1
 
 π  6⋅Φ  3  ζ 3  Φ 3 2

XfNJ (Φ ,N p ,Φ max ): = α ⋅ ⋅  ⋅ N 2 ⋅  2  ⋅ 1–   + (1 – Φ 3 )
8  π ⋅ N p    Φ max  
p
   
 

E10: = 0.2

G’ data on butyl rubber/N330 cpds G’ data on butyl rubber/N330 cpds


10 10
Modulus, MPa
Modulus, MPa

5 5

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
Carbon black volume fraction Carbon black volume fraction
Guth & Gold Data Network junction Data

E20 : = 5.625
E’ data on SBR 1500/N347 cpds E’ data on SBR 1500/N347 cpds
40 40
Modulus, MPa

Modulus, MPa

20 20

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
Carbon black volume fraction Carbon black volume fraction
Guth & Gold Data Network junction Data
196 Filled Polymers

E20 : = 5.625
E’ data on SBR 1500/N347 cpds E’ data on SBR 1500/N347 cpds
40 40
Modulus, MPa

Modulus, MPa
20 20

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
Carbon black volume fraction Carbon black volume fraction
Guth & Gold Data Network junction Data

A5.2 Kraus Deagglomeration–Reagglomeration Model


for Dynamic Strain Softening
[G. Kraus. J. Appl. Polym. Sci.: Appl. Symp., 39, 75–92, 1984.]

A5.2.1  Soft Spheres Interactions

Aggregates are viewed as “soft spheres” in such


a manner that, over a range of distance rm – rM,
the interparticle force increases with the dis-
rm
placement from the equilibrium distance rm

Dynamic modulus varies with strain because


of a deagglomeration–reagglomeration process
of soft spheres. d

 = >  difference between the two limiting elas-


tic modulus values:
rm rM
   (Ge0 –Geinf ) = C1 ⋅ N 0 ⋅ kB ⋅ (d + rm )2

C1 is a constant   Ge0 : limiting low strain


modulus ∂F
        Geinf : limiting high strain ∂r r = r
Force F(r)

m
modulus 0

∂F
kB =  = rate constant of the deagglomera-
∂r r = rm tion process

Note: G′ = Ge; G″ = Gv Interparticle distance r


Polymers and Carbon Black 197

A5.2.2  Modeling G′ vs. γ0


A power law model is assumed to suit the strain dependency of the deagglom-
eration and agglomeration processes
deagglomeration:   γ0m   reagglomeration :γ0−m
Rem: this means that both processes are perfectly symmetrical, as illustrated
below:

γ0: =  0,0.01..5   m: = 0.5   f ( γ 0 ): = γ 0m    g( γ 0 ): = γ −0 m

10

< = deagglomeration
1
< = reagglomeration

0.1
0 1 2 3 4 5
γ0

f(γ0) g(γ0) 1

At equilibrium the rates of aggregate deagglomeration and reagglomera-


tion are equal, i.e.:

kB ⋅ N ⋅ γ m0 = kR ⋅ ( N 0 − N ) ⋅ γ 0− m kB: rate constant of deagglomeration


kR: rate constant of reagglomeration
N0: number of contacts at zero strain
N: number of surviving contacts
kR ⋅ γ −0 m ⋅ N 0 γ −0 m ⋅ N 0
=> N =     or:   N =
kB ⋅ γ m0 + kR ⋅ γ −0 m kB m k R − m
⋅γ0 + ⋅γ0
kR kR

One defines a critical strain that corresponds 2⋅m


k
to the deagglomeration–reagglomeration equilibrium:   γ c =  R 
 kB 

N0 With respect to the attraction-repulsion


It follows:  N = 2⋅m mechanism between soft spheres (aggre-
γ  gates), there is an excess modulus due to
1+  0 
 γc  the agglomerate network at any ampli-
tude that must be proportional to N, i.e.:
(Ge(γ0)−Geinf) ⋅ N0 = (Ge0−Geinf) ⋅ N
Note: here G* = Gx; G’ = Ge; G″ = Gv
198 Filled Polymers

Ge( γ 0 ) – Geinf 1 (Ge0 − Geinf )


=> = 2⋅m
or:   Ge( γ 0 ) = Geinf +
Ge0 – Geinf γ    γ 0  2⋅m 
1+  0  1 +   
 γc    γ c  

Numerical illustration:

γ0 : = 0.001,0.002,..10   : strain range for calculation

Geinf: = 2   Ge0: = 22   m: = 0.55   γc: = 0.03   : model parameters

100
Ge0 − Geinf
Ge( γ 0 ): = Geinf + 2⋅m γc
γ 
1+  0 
 γc 
10
Ge0 + Geinf
= 12
2

Ge( γ c ) = 12
1
1.10–3 0.01 0.1 1 10
γ0
Ge0+Geinf
Ge(γ0) 2

There is indeed horizontal symmetry with


Ge0 + Geinf
respect to the mid modulus at
the critical strain γ . 2
c

A5.2.3 Modeling G″ vs. γ0


There is an excess loss modulus that may be taken as proportional to the rate
of deagglomeration (or reagglomeration), i.e.:

Gv( γ ) − Gv inf = c ⋅ kB ⋅ γ 0m ⋅ N [1] Gvinf: limiting high strain viscous


modulus
Rem: Kraus gives no indication
regarding the values of Gvinf except
that Gvinf << Gm.

One has (at deagglomeration–agglomeration N0


equilibrium): N= 2⋅m
   [2]
γ 
1+  0 
 γc 
Polymers and Carbon Black 199

From the soft spheres approach, one has:

Ge0−Geinf = C1⋅N0⋅kB⋅(d + rm)2    where C1 is a constant

Ge0 − Geinf
It follows:    N 0 ⋅ kB = [3]
c ⋅ (d + rm )2

Equations [2] and [3] are sub-


C1 Ge0 − Geinf
stituted in Equation [1], then:   Gv( γ ) − Gv inf = ⋅ γ m0 ⋅
[ c ⋅ (d + rm )2 ] γ 
2⋅m

1+  0 
 γc 

C1
All terms in are constant, then:
[ + rm )2 ]
c ⋅ ( d

C ⋅ γ m ⋅ (Ge0 − Geinf ) C1
Gv( γ ) = Gv inf +     with C = 
γ 
2⋅m
[ c ⋅ (d + rm )2 ]
1+  0 
 γc 

C ⋅ γ m ⋅ (Ge0 − Geinf )
The function Gv( γ ) = Gv inf + 2⋅m
has a maximum value Gvm
for γ0 = γc γ 
1+  0 
 γc 

C ⋅ γ mc ⋅ (Ge0 − Geinf ) which gives for the constant: C = 2 ⋅ (Gv m − Gv inf )


Gv m = Gv inf +
γ 
2 ⋅m
[ γ mc ⋅ (Ge0 − Geinf )]
1+  c 
 γc 

It follows:


( γγ )
m
( Gv m − Gv inf ) ( Ge0 − Geinf ) ( Gv m − Gv inf )
Gv ( γ ) = Gv inf + 2 ⋅ ⋅γm ⋅ Gv ( γ ) − Gv inf = 2 ⋅ 0
[ γ c m ⋅ ( Ge0 − Geinf )]
( ) ( )
2⋅m 2⋅m
 γ0  c  γ0 
1 + γ  1 + γ 
 c   c 

m   γ0 m
γ  2 ⋅   
2 ⋅ 0  or:
Gv( γ 0 ) − Gv inf  γc    γ c  
=> = Gv( γ 0 ) = Gv inf + (Gv m − Gv inf ) ⋅ 
( Gv m − Gv inf )   γ 0  2⋅m    γ 0  2⋅m 
1 +    1 +   
  γ c     γ c  
200 Filled Polymers

Numerical illustration:

γ0 : = 0.0001, 0.002,..10   : strain range for calculation

Gvinf: = 0.001   Gvm: = 1.6   m: = 0.55   γc: = 0.03   : model parameters

  γ0 m
2 ⋅   
  γ c  
Gv(γ 0 ) : = Gv inf + (Gv m − Gv inf ) ⋅ 
  γ 0  2⋅m 
1 +   
  γ c  

Gv(γc) = 1.6 10
γc

There is indeed vertical symmetry


with respect to the critical strain γc. 0.1
1.10–4 1.10–3 0.01 0.1 1 10
γ0
Gv(γ0) Gvm

lim Gv ( γ 0 ) → 1.0000000000000000000.10−3 = 1 ⋅ 10−3


γ 0 →0

lim Gv( γ 0 ) → 1.0000000000000000000.10−3 = 1 ⋅ 10−3


γ 0 →∞

Rem: the low strain limiting value of the viscous


modulus Gv is obviously the value assigned to
Gvinf; in agreement with Kraus paper, one has con-
sidered Gvinf << Gvm

A5.2.4  Modeling tan δ vs. γ0


One has thus:

( ) 
m
 γ0
   and   2 ⋅ γ
(Ge0 − Geinf ) 
Ge( γ 0 ) = Geinf + Gv ( γ 0 ) = Gv inf + ( Gv m − Gv inf ) ⋅ c

  γ 0  2⋅m 
( ) 
2 ⋅m
 γ0
1 +    1 + γ
  γ c    c
Polymers and Carbon Black 201

Gv( γ 0 )
By definition: tan δ =
Ge( γ 0 )

γ0
Let: X =
γc


 [ Gv inf ⋅ ( 1 + X ) + ( Gv m − Gv inf ) ⋅ 2 ⋅ X ] 
2⋅m m

 ( 1 + X 2⋅m )  [ Gv inf ⋅ ( 1 + X 2⋅m ) + ( Gv m − Gv inf ) ⋅ 2 ⋅ X m ]


tan δ = =
[ Geinf ⋅ ( 1 + X 2⋅m ) + ( Ge0 − Geinf )] [ Geinf ⋅ ( 1 + X 2⋅m ) + ( Ge0 − Geinf )]
(1 + X )2 ⋅m


tan δ =
( Gvinf + Gvinf ⋅ X 2⋅m + 2 ⋅ X m Gvm − 2 ⋅ X m ⋅ Gvinf ) = X m ⋅ ( 2 ⋅ Gvm + Gvinf ⋅ X m − 2 ⋅ Gvinf + Gvinf ⋅ X − m )
( Geinf ⋅ X 2⋅m + Ge0 ) X m ⋅ ( Geinf ⋅ X m + Ge0 ⋅ X − m )


2 ⋅ Gv m + Gv inf ⋅ X m − 2 ⋅ Gv inf + Gv inf ⋅ X − m 2 ⋅ Gv m + Gv inf ⋅ ( X m − 2 + X − m )
tan δ = =
Geinf ⋅ X m + Ge0 ⋅ X − m ( Geinf ⋅ X m + Ge0 ⋅ X − m )

m −m
m −m
2 ⋅ Gv m + Gv inf ⋅ (X 2 –X 2 )2 m − 2 + X − m ) = ( X 2 − X 2 )2
or  tan δ =   since  ( X
Geinf ⋅ X m + Ge0 ⋅ X − m

−m 2
 m

  γ0  2  γ0  2 
Gv inf ⋅   −   + 2 ⋅ Gv m
 γ c   γc  
=> tan δ( γ 0 ) =  
m −m
γ
  γ 
Geinf ⋅  0  + Ge0 ⋅  0 
 γc   γc 

γ0: = 0.001, 0.002..100

Gvinf: = 0.001    m: = 0.55    γc: = 0.03    Gvm: = 1.6

−m 2
 m

  γ0  2  γ0  2 
Gv inf ⋅   −   + 2 ⋅ Gv m
 γ c   γc  
tanδ( γ 0 ) : =      tanδ(γc) = 0.133
m −m
 γ0   γ0 
Geinf ⋅   + Ge0 ⋅  
 γc   γc 
202 Filled Polymers

1
Gv m γc
= 0.133
Ge( γ c )
0.1

The tan δ curve exhibits also a


0.01
(vertical) symmetry but not with 1.10–3 0.01 0.1 1 10 100
respect to γc. γ0
Gvm
tanδ (γ0)
Ge (γc)

A5.2.5  Complex Modulus G* vs. γ0


One has:  Gx( γ 0 ) = Ge( γ 0 )2 + Gv( γ 0 )2   Rem: absolute value of the c­ omplex modulus

 [ Geinf ⋅ (1 + X 2⋅m ) + (Ge0 − Geinf )]   [ Gv inf ⋅ (1 + X 2⋅m ) + (Gv m − Gv inf ) ⋅ 2 ⋅ X m ] 


2 2
  ⇒ Gx( γ 0 ) =   + 
 (1 + X 2⋅m )   (1 + X 2⋅m ) 

γ0
with X =
γc

1
Gx ( γ 0 ) = ⋅ (Geinf ⋅ X 2⋅m + Ge0 )2 + (2 ⋅ Gv m ⋅ X m + Gv inf + Gv inf ⋅ X 2⋅m − 2 ⋅ Gv inf ⋅ X m )2
(1 + X 2⋅m )

this term is suitably rearranged as:

2 ⋅ Gv m ⋅ X m + Gv inf ⋅ X m ⋅ (X − m + X m − 2)
m −m
2 ⋅ Gv m ⋅ X m + Gv inf ⋅ X m ⋅ (X 2 − X )
2 2

−m
 m

X m ⋅  2 ⋅ Gv m + Gv inf ⋅ (X 2 − X 2 )2 
 

−m 2
1   m

  => Gx( γ 0 ) = ⋅ (Geinf ⋅ X 2⋅m + Ge0 )2 +  X m ⋅  2 ⋅ Gv m + Gv inf ⋅ (X 2 − X 2 )2  
1+ X 2 ⋅m
  

i.e.:
2
2⋅m 2    m −m 2 

1   γ0    γ 0  m   γ 0  2  γ 0  2  
Gx(γ 0 ) = 
⋅  Geinf ⋅   + Ge0  +   ⋅  2 ⋅ Gv m + Gv inf ⋅   −  
2⋅m
  γ 0     γc    γ c    γ c   γ c   
1 +         
  γ c  
Polymers and Carbon Black 203

Alternatively one can consider:

  Gx = Ge2 + Gv 2 => Gx = Ge2 + tanδ 2 ⋅ Ge2 => Gx = Ge ⋅ 1 + tanδ 2

Substitution gives:
2
  m −m 2
 
     γ  2  γ  2
 
 Gv inf ⋅  γ  −  γ   + 2 ⋅ Gv m 
0 0
 
 (Ge0 − Geinf )   
c c
 
Gx( γ 0 ) =  Geinf + 2⋅m 
⋅ 1+  
  γ0     γ0 
m
 γ0 
–m

 1 +     Geinf ⋅   + Ge0 ⋅  
   γ      γ   γ  
  c  
c c

 

Numeric illustration
γ0: =  0.001, 0.002..10

Ge0 = 22   Geinf = 2   m = 0.55   γc = 0.03   Gvm = 1.6   Gvinf = 1.10−3

2
2⋅m 2    m −m

2

  γ0    γ 0  m 
Gx( γ 0 ) : =
1
⋅  Ge ⋅ + Ge  + ⋅ 2 ⋅ Gv + G v ⋅  γ 0  2 γ 
– 0 
2
 
inf 
  γ 0  2⋅m    γ c  0   
  γ c  
m inf
 γ c   γc   
1 +      
  
  γ c  

  lim Gv( γ 0 ) → 1.0000000000000000000.10−3 < = one recovers obviously


γ 0 →∞
the value assigned to Gvinf
Gv0: = Gvinf   Gvinf = 1.10−3
100
Gx 0 : = Ge0 2 + Gv 0 2 γc

Gx0 = 22   < = obviously = Ge0

10
Gx inf : = Geinf 2 + Gv inf 2

Gxinf = 2   < = obviously = Geinf

1
1.10–3 0.01 0.1 1 10
γ0
Gx(γ0) Ge(γ0) Gx(γc)
Ge0 Geinf
204 Filled Polymers

Gx 0 + Gx inf
Mid modulus: = 12
2

The mid complex modulus G* does not coincide with G* at γc  Gx(γc) = 12.106

A5.2.6 A Few Mathematical Aspects of the Kraus Model


One has thus:
m
γ 
2 ⋅ 0 
Ge( γ 0 ) − Geinf 1 Gv( γ 0 ) − Gv inf  γc 
=    and   =
Ge0 − Geinf γ 
2⋅m
(Gv m − Gv inf ) γ 
2⋅m

1+  0  1+  0 
 γc   γc 

Let:
m
γ 
2 ⋅ 0 
1  γc 
Z1( γ 0 ) : =     Z 2( γ 0 ) : =
  γ 0  2⋅m    γ 0  2⋅m 
1 +    1 +   
  γ c     γ c  

lim Z1( γ 0 ) → 1.     lim Z2( γ 0 ) → 0


γ 0 →0 γ 0 →0

lim Z1( γ 0 ) → 0     lim Z2( γ 0 ) → 0


γ 0 →∞ γ 0 →∞

1
γc

0.5 The actual shapes of G*(γ0), G′(γ0)


and G″(γ0) are essentially dictated by
the mathematical functions used to
express the dependence on γ0 and by
0 the upper and lower limits.
1.10–3 0.01 0.1 1 10
γ0
Z1(γ0) Z2(γ0)
Polymers and Carbon Black 205

Numerical illustration
γ0: = 0.0001, 0.0002..10
Ge0: = 15.106⋅Pa   Geinf: = 3⋅106 ⋅ Pa  Data: G. Heinrich, M. Kluepel. Adv. Polym. Sci., 160,
1–44 , 2002; Figure 1; NR + N110 cpd, vulcanized.

Gvm: = 1.7 ⋅ 10 ⋅Pa  6 Gvinf: =  0.1 ⋅ 10 ⋅Pa 6

m: =  0.55     γc: = 0.013

  Gx = Ge ⋅ 1 + tanδ 2    Gx = A ⋅ 1 + B   The 1 + tan δ 2 term can


be viewed as the viscous
character, weighed with
respect to elasticity.
Let:
2
  m −m 2
 
   γ0  2  γ0  2  
 Gv inf ⋅  γ  −  γ   + 2 ⋅ Gv m 
 
c c
 
 (Ge0 − Geinf )     B( γ 0 ) : =  
A( γ 0 ) : =  + Geinf    γ0 
m
 γ0 
–m

  γ 0  ( 2⋅m)  Ge ⋅
inf  + Ge ⋅
 1 +     γ c  0 
 γ c  
   γ c     
   

=> Gx( γ 0 ) := A( γ 0 ) ⋅ 1 + B( γ 0 )

2.107
γc
A(γ0) is the elastic modulus and the
variation of both G′ ( = Ge) and G*
( = Gx) with strain are similar.
1.107

γc
1.10–4 1.10–3 0.01 0.1 1 10 1.04
γ0
A(γ0) Gx(γ0) Gx(γc)
1.02

1 + B( γ c ) = 1.018 1.10–4 1.10–3 0.01 0.1 1 10


γ0

1+B(γ0) 1+B(γc)
206 Filled Polymers

The Kraus model implies thus that, at the critical strain, the viscous char-
acter is maximum 2% of the elastic one and less than 0.1% at very high strain.
Such limits somewhat depend on the value of the critical strain γc.
The variations of G′ and G″, as modeled by Kraus model, essentially reflect
the mathematical forms of the γ0/γc functions.

m
γ 
2 ⋅ 0 
1     γc     Zx( γ 0 ): = Z1(γ 0 )2 + Z 2(γ 0 )2
Z1( γ 0 ) : = Z 2( γ 0 ) :=
  γ 0  2⋅m    γ 0  2⋅m 
1 +    1 +   
  γ c     γ c  

ZX(γc) = 1.118

1
γc
γc

0.5
1

0 0.8
1.10–4 1.10–3 0.01 0.1 1 10 1.10–4 1.10–3 0.01 0.1 1 10
γ0 γ0
Z1(γ0) Z2(γ0) Zx(γ0) Zx(γ0)
   

A5.2.7  Fitting Model to Experimental Data


DSS data on SBR/60 phr black cpds
[M. Gerspacher, C.P. O’Farrell, C. Tricot, L. Nikiel, H.A. Yang. ACS Rubb.
Div; Mtg, Louisiana, KY, 1996. Paper 74.]
Polymers and Carbon Black 207

Strain sweep experiments

N660 N330 N110 Kraus model


Strain G ′ G ′′ G′ G ′′ G′ G ′′ Elastic modulus
% MPa MPa MPa MPa MPa MPa
(Ge0 − Geinf )
Ge(γ 0 ) = Geinf +
  γ 0  2 ⋅m 
 0.091 2.34 0.265 4.45 0.630 6.02 0.915  1 +   
    γ c  
 0.19 2.31 0.275 4.41 0.630 5.95 0.938 
 0.28 2.3
30 0.280 4.30 0.650 5.80 0.942 
 
 0.38 2.26 0.290 4.24 0.660 5.61 0.960  Geinf = G′inf: elastic modulus
 0.48 2.24 0.295 4.10 0.670 5.50 980 
0.9 for an infinite strain
 
 0.55 2.23 0.296 4.05 0.685 5.32 0.993  Ge0 = G′0: initial low strain
 
 0.66 2.22 0.297 3.95 0.692 5.18 1.000  elastic modulus
 0.74 2.21 0.297 3.89 0.7
700 5.05 1.010  γc: critical strain
 
 0.84 2.20 0.297 3.82 0.710 4.95 1.026 
 0.93
 2.18 0.298 3.75 0.716 4.83 1.030  Viscous modulus
 1.07 2.16 0.2298 3.61 0.720 4.65 1.032 
    γ0 m
Data: =  1.22 2.15 0.299 3.55 0.720 4.48 1.033  2 ⋅  
 1.48 1.031    γ c  

2.14 0.299 3.46 0.719 4.31
 Gv(γ 0 ) = Gv inf + (Gv m − Gv inf ) ⋅ 
 1.7 2.12 0.298 3.40 0.718 4.20 1.030    γ 0  2 ⋅m 
  1 +   
 1.9 2.11 0.298 3..34 0.717 4.08 1.020    γ c  
 2.8 2.06 0.294 3.08 0.690 3.71 0.968 
 
 3.7 2.05 0.290 2.95 0.660 3.49 0.914  Gvinf = G″inf: viscous modulus
 4.6 0.870 

2.04 0.294 2.85 0.631 3.30
 at infinitely low or infinitely
 5.5 2.02 0.285 2.78 0.608 3.18 0.828  high strain
 
 6.5 2.00 0.280 2.71 0.581 3.08 0.782 
Gvm = G″m: maximum value
 7.6 1.99 0.275 2.65 0.560 3.01 0.752 
  of viscous modulus (occur-
 8.5 1.97 0.271 2..62 0.542 2.95 0.722 
 0.700 
ing at the critical strain)
 9.5 1.95 0.269 2.57 0.530 2.90
m = exponent related to the
rates of deagglomeration
and reagglomeration of filler
particles
γ0 = strain amplitude

A5.2.7.1  Modeling G′ vs. Strain


Extracting data   Strain (%)   N660 G′ (MPa)   N330G′ (MPa)   N110G′ (MPa)

j: = 0..2    γ: = Data<0>    Ge0: = Data<1>    Ge1: = Data<3>    Ge2: = Data<5>

n: = length(γ)   n = 23   :number of data

Guess parameters for nonlinear fitting algorithm (GenFit function)


208 Filled Polymers

 (Ge j )n−1  < = extracting guess parameters from experimental data, i.e.,


  - The lowest measured modulus for Ginf′
 (Ge j )0 
Cj : = γ  - The highest measured modulus for G0′
 round n   - The mid range strain for γc
 2
 - The common value considered by Kraus
 0.6 

 C1 − C0 
 C0 + 2⋅C3 
  γ  
1+  
  C2  
 
 1 
1−
   γ  2⋅C3  
 1+    
   C2   
 
 1 
   γ  2⋅C3  
Model equation and  1+    
partial derivatives:    F(γ ,C): =    C2   
 2⋅C3 
 C1 − C0  γ  C3 
 2⋅ ⋅  ⋅ 
2⋅C3 2  C  C2
   γ   2

 1 +    
  C  
 2

   
   
  −  γ
2⋅C3
  γ 
− 2 ⋅ C C   
1 0
⋅    ⋅ ln  
    γ  2⋅C3  2  C2    C2  
  1 +     
    C2    

N660 N330 N110

 1.95   2.57   2.9  :initial guessed parameters


 2.34   4.45  6.02  as extracted from the exper-
C0 =   C1 =   C2 =  
imental data sets
 1.48   1.48   1.48 
     
 0.6   0.6   0.6 

Resj: = GenFit(γ,Gej,Cj,F) : calling the nonlinear regression algorithm


Polymers and Carbon Black 209

Fitting equation Correlation coefficient r2


 
    
    
 (Res j )1 − (Res j )0   (Res j )1 − (Res j )0  
Ga j : = (Res j )0 + 2⋅( Res j )3 
Gar2 j : = corr  (Res j )0 + 2⋅( Res j )3 
, Ge j 
  γ     γ   
 1 +     1+    
  ( Res j ) 2      ( Res j ) 2   

Result: fit parameters

   N660 N330 N110

 1.825   2.363   2.621 ′


:Ginf
 2.414   4.581  6.209 
Res 0 =      Res1 =      Res 2 =      : G0′
 1.761   1.471   1.376  : γc
     
 0 . 344   0 . 573   0 . 614  :m

   Gar20 = 0.998 Gar21 = 0.999 Gar22 = 1 : r2

8
Elastic modulus, MPa

0
0.01 0.1 1 10 100
Strain, %
Fit N660 Data N660 Fit N330
Data N330 Fit N110 Data N110

A5.2.7.2  Modeling G″ vs. Strain


Extracting data  Strain(%)  N660G″ (MPa)  N330G″ (MPa)  N110G″ (MPa)

J: = 0..2      γ: = Data<0>  Gv0: = Data<2>    Gv1: = Data<4>    Gv2: = Data<6>

Guess parameters for nonlinear fitting algorithm (GenFit function)


Finding max G″ and corresponding γ
210 Filled Polymers

gc j : = A ← max ( Gv j ) < = extracting guess parame-


for i ∈1.. n − 1
(Gv j )n −1  ters from experimental data,
  i.e.:
   (gc j )1 
 γi  Cj : = 
B←  if (Gv j )i = A (gc j )0  - The measured modulus
 ( Gv )   
j i
 0.5 
′′
at the lowest strain for Ginf
A←B - The maximum ­measured
A modulus for Gm ′′
- The strain for the maxi-
mum G”
- The common value con-
sidered by Kraus

Model equation and partial derivatives:

  γ  3
C

 2 ⋅  
 C2 
 C0 + ( C1 − C0 ) ⋅ 
  γ 
2⋅C3 
 1+   
  C2  
 C3 
   γ 
2 ⋅ 
  C2  
 1− 2⋅C3 
  γ  
1+  
  C2  
 C3

  γ  
 2 ⋅  
 C2 
   Q(γ ,C): =  2⋅C3

  γ  
 1+   
  C 2  
 C3  2⋅C3  
  γ    γ   
  C   C 
C3   
 2 ⋅ ( C0 − C1 ) ⋅ ⋅ ⋅ 2⋅ − 1 
2 2

 C2   γ  2⋅C3     γ  2⋅C3 
1 +     1 +     
  C    C   
  2    2  
  2⋅C3  
  γ    γ   
C ln    C 
  γ  3  C2   
 2 ⋅ ( C1 − C0 ) ⋅   ⋅ ⋅ 1− 2⋅ 2

 C2   γ  2⋅C3     γ  2⋅C3   
1 +     1 +     
   C2      C2    

Polymers and Carbon Black 211

N660 N330 N110

 0.269   0.53   0.7  : initial guessed parameters


 0.299   0.72  1.033  as extracted from the experi-
C0 =   C1 =   C2 =  
 1.48   1.22   1.22  mental data sets
     
 0.5   0.5   0.5 

Resj: = GenFit(γ,Gvj,Cj,Q) : calling the nonlinear regression algorithm

Fitting equation


  γ 
( Res j )3

 2⋅  
  (Res j )2  
 
Gb j := (Res j )0 + (Res j )1 − (Res j )0  ⋅ 2⋅( Res j )3 
  γ  
 1+   
  ( Res j ) 2  

Correlation coefficient r2


 ( Res j )3
 
  γ  
2⋅  
  ( R es )
j 2   
Gbr2 j := corr  (Res j )0 + (Res j )1 − (Res j )0  ⋅ 2⋅( Res j )3 
, Gv j 
  γ   
 1+    
   (Res j )2   

    N660 N330 N110

 0.224   0.477   0.659  ′′


:Ginf
 0.3   0.713   1.026 
    : Gm ′′
Res 0 =      Res1 =      Res 2 = 
 2.174   1.808   1.428  : γc
     
 0.491  0.706   0.547  :m

    Gbr20 = 0.709 Gbr21 = 0.64 Gbr22 = 0.692 : r2


212 Filled Polymers

The Kraus model fits reasonnably


1 well the experimental G′ data but
Viscous modulus, MPa

cannot meet the asymmetric shape


of G” vs. strain data.

0.5
Note: the results of the nonlinear
algorithm are very sensitive to the
initial guess values, particularly for
0 γc; using the strain position of Gmax
′′
0.01 0.1 1 10 100
Strain, %
somewhat reduces this problem.

A5.3 Ulmer Modification of the Kraus Model for Dynamic


Strain Softening: Fitting the Model
Elastic modulus

Ge( γ 0 ) = Geinf +
( Ge0 − Geinf )
  γ 0  2⋅m 
1 +   
  γ c  

Viscous modulus (Ulmer’s modification)

m
γ 
2 ⋅ ( Gv m − Gv inf ) ⋅  0 
 γc   −γ 
Gv( γ 0 ) = Gv inf + + Gv k ⋅ exp  0 
γ 
2⋅m
 γk 
1+  0 
 γc 

Dynamic strain softening data on SBR/60 phr black cpds


[M. Gerspacher, C.P. O’Farrell, C. Tricot, L. Nikiel, H.A. Yang. ACS Rubb.
Div; Mtg, Louisiana, KY, 1996. Paper 74]
Polymers and Carbon Black 213

Strain sweep experiments


Geinf = G’inf: elastic modu-
N660 N 330 N110 lus for an infinite strain
Strain G′ G ′′ G′ G ′′ G′ G ′′
% MPa MPa MPa MPa MPa MPa Ge0 = G’0: initial low
strain elastic modulus
0.091 2.34 0.265 4.45 0.630 6.02 0.915 

 0.19 2.31 0.275 4.41 0.630 5.95 0.938 
Gvinf = G″inf: viscous mod-
 0.28 2.330 0.280 4.30 0.650 5.80 0.942 
  ulus at infinitely low or
 0.38 2.26 0.290 4.24 0.660 5.61 0.960 
high strain
 0.48 2.24 0.295 4.10 0.670 5.50 0.9980 
 
 0.55 2.23 0.296 4.05 0.685 5.32 0.993 
  Gvm = G ″m: maximum
 0.66 2.22 0.297 3.95 0.692 5.18 1.000 
value of viscous modu-
 0.74 2.21 0.297 3.89 0.7700 5.05 1.010 
  lus (occuring at the criti-
 0.84 2.20 0.297 3.82 0.710 4.95 1.026 
 0.93 cal strain)
2.18 0.298 3.75 0.716 4.83 1.030 
 
 1.07 2.16 0.2298 3.61 0.720 4.65 1.032 
  Gvk = G ″k drop in viscous
Data :=  1.22 2.15 0.299 3.55 0.720 4.48 1.033 
 1.48
modulus assumed to be
2.14 0.299 3.46 0.719 4.31 1.031 
  proportional to the num-
 1.7 2.12 0.298 3.40 0.718 4.20 1.030 
 1.9 ber of contacts between
2.11 0.298 3..34 0.717 4.08 1.020 
  aggregates at zero strain
 2.8 2.06 0.294 3.08 0.690 3.71 0.968 
 
 3.7 2.05 0.290 2.95 0.660 3.49 0.914 
m = exponent related to
 4.6 2.04 0.294 2.85 0.631 3.30 0.870 
  the rates of deagglomer-
 5.5 2.02 0.285 2.78 0.608 3.18 0.828 
 6.5 ation and reagglomera-
2.00 0.280 2.71 0.581 3.08 0.782 
  tion of filler particles
 7.6 1.99 0.275 2.65 0.560 3.01 0.752 
 
 8.5 1.97 0.271 2..62 0.542 2.95 0.722 
γ0 = strain amplitude
 9.5 1.95 0.269 2.57 0.530 2.90 0.700 

γc = critical strain for G ″m

γk = critical strain for G ″k

A5.3.1  Modeling G′ vs. Strain (same as Kraus)


Extracting data  Strain (%)  N660 G′ (MPa)  N330 G′ (MPa)  N110 G′ (MPa)
j : = 0..2      γ: = Data<0>  Ge0: = Data<1>   Ge1: = Data<3>   Ge2: = Data<5>
n: = length(γ)   n = 23   : number of data
Guess parameters for nonlinear fitting algorithm (GenFit function)
214 Filled Polymers

 ( Ge j )n −1  < = extracting guess parameters from experimental


  data, i.e.,
 ( Ge j )0  - The lowest measured modulus for G ′inf
Cj : =   - The highest measured modulus for G’0
 γ round  n  
  2   - The mid range strain for γc
  - The common value considered by Kraus
 0.6 

 C1 − C0 
 C0 + 2⋅C3 
  γ  
1+  2 
 C  
 
 1 
1−
   γ  2⋅C3  
 1 +    
   C2   
 
 1 
   γ   2⋅C3

Model equation and  1 +    
partial derivatives: F(γ ,C) : =    C2   
 2⋅C3 
 C1 − C0  γ  C3 
 2⋅ ⋅  ⋅ 
2⋅C3 2  C  C
   γ   2 2

 1 +    
  C2  
 
   
   
  C1 − C0  γ  3  γ 
2⋅C
− 2 ⋅ ⋅    ⋅ ln   
    γ  2⋅C3  2  C2    C2  
  1 +     
    C2    

N660 N 330 N110 :initial guessed parameters as


extracted from the ­experimental
 1.95   2.57   2.9  data sets
 2.34   4.45  6.02 
C0 =   C1 =   C2 =  
 1.48   1.48   1.48 
     
 0.6   0.6   0.6 

Resj: = GenFit(γ, Gej, Cj, F) : calling the nonlinear regression algorithm


Polymers and Carbon Black 215

Fitting equation Correlation coefficient r2


 
    
     
 (Re s j )1 − (Re s j )0   (Re s j )1 − (Re s j )0  
Ga j : = ( Re s j )0 + 2⋅(Re s j )3 
Gar2 j : = corr  ( Re s j )0 + 2⋅(Re s j )3 
, Ge j 
  γ     γ   
 1+     1+    
  (Re s j )2      (Re s j )2   

Results: fit parameters

N660 N 330 N110

 1.825   2.363   2.621 : Gi′nf


 2.414   4.581  6.209  : G0
Res 0 =   Res1 =   Res 2 =  
 1.761   1.471   1.376  : γc
     
 0.344   0.573   0.614  :m

Gar 2 0 = 0.998 Gar 2 1 = 0.999 Gar 2 2 = 1 :r 2

8
Elastic modulus, MPa

0
0.01 0.1 1 10 100
Strain, %
Fit N660 Data N660 Fit N330
Data N330 Fit N110 Data N110

A5.3.2  Modeling G′′ vs. Strain

Extracting data   Strain (%)   N660 G″ (MPa)   N330 G″ (MPa)   N110 G″ (MPa)

j: = 0..2      γ : = Data<0>  Gv0 : = Data<2>   Gv1 : = Data<4>   Gv2: = Data<6>

Guess parameters for nonlinear fitting algorithm (GenFit function)


216 Filled Polymers

Finding max G′′ and corresponding γ

gc j := A ← max ( Gv j )
for i ∈1..n − 1
 γi 
B←  if ( Gv j )i = A
( Gv j )i 
A←B
A
  γ  3
C

 2 ⋅ (C1 − C0 ) ⋅   
C +  C2   −γ  
+ C ⋅ exp  C  
(Gv j )n −1   0  γ 
2⋅C3 4
5
   1+   
( gc )   C2  
 j 1 
 
 (gc )    γ 
C3

 j 0
   C  
       C j :=  0.55   1− 2⋅ 2
2⋅C3 
   γ   
  1 +
   
 (Gv j )0  
   C 2  


 (gc )   C3 
 j 0    γ  
 100   
 C2   
 2⋅ 
  γ  2⋅C3 
 1 +    
   Model equation and    C2   
 
partial derivatives: Q( γ , C) := 
 γ  3  γ  3 
C 3⋅C

   −  C  
 2 ⋅ (C − C ) ⋅ C3 ⋅  C2  2 
Extracting guess parameters  0 1
C2 2⋅C3 2 
   γ   
from experimental data, i.e.:   1+   


  C 2  

 
  γ  3  γ  3 
3⋅C
- The measured modulus at the C
   −  
highest strain for G″inf  2 ⋅ (C − C ) ⋅ ln  γ  ⋅  C2   C2  
- The maximum measured  0 1  C  2⋅C 2 
 2   γ  3 
modulus for G m ″
  1 +    
   C2   
- The strain for the maximum
 
G″  exp  
 −γ  
- The common value consid-   C5  
 
ered by Kraus  γ  −γ  
C ⋅ ⋅ exp
- The difference G ″m –G″0 (mea-  4
( C ) 2
 C5  
 5 
sured modulus at the lowest
strain) for G″k
- γc/100 for γk
Polymers and Carbon Black 217

N660 N 330 N110

 0.269   0.53   0.7 


 0.299   0.72   1.033 
      : initial guessed parame-
 1.48   1.22   1.22  ters as extracted from the
C0 =   C1 =   C2 =   experimental data sets
 0.55   0.55   0.55 
 0.265   0.63   0.915 
     
 0.015   0.012   0.012 

Resj: = GenFit (γ, Gvj, Cj, Q) : calling the nonlinear regression algorithm

Fitting equation:


 γ 
(Re s j )3

 2⋅  
    (Re s j )2    −γ 
Gb j : =  (Re s j )0 + (Re s j )1 − (Re s j )0  . 2⋅(Re s j )3
+ (Re s j )4 ⋅ exp  
  γ    (Re s j )5    
 1 +   
   (Re s j )2   

r2:

  
 (Re s j )3 
 γ  
 2⋅  
    ( Re s j )2    −γ  

( 0
)
Gbr2 j : = corr   Re s j + ( Re s j )1 − ( Re s j )0  .
 
2⋅( Re s j )
+ ( Re s j )4 ⋅ exp



(
 Re s )





, Gv j 

 γ  3
  j 5 
 1+    

   (
 Re s j 
2
) 


N660 N 330 N110

 − 0.284   0.087   0.12  ′′


: Ginf
 0.3   0.723   1.036 
      : Gm′′
 1.271   1.351   1.295  : γc
Res 0 =   Res1 =   Res 2 =  
 0.162   0.466   0.519  :m
 0.13   0.371   0.572  : Gk′′
     
 0.042   0.156   0.199  : γk
Gbr 2 0 = 0.985 Gbr 2 1 = 0.999 Gbr 2 2 = 1 : r2
218 Filled Polymers

Viscous modulus, MPa


0.5

0
0.01 0.1 1 10 100
Strain, %

The Ulmer’s additional term to the Kraus equation for G ″ vs. strain gives a
considerably improved fit of experimental data.

A5.4 Aggregates Flocculation/Entanglement Model


(Cluster–Cluster Aggregation Model, Klüppel et al.)
Aggregate solid fraction

3− F
 d  F = fractal dimension
φ A (D) =  
 D of aggregate ( = 1.8) a
D

Elastic modulus of aggregate:

 3+ FB 
 
GA = Gp ⋅ φA  3− F 

FB = fractal dimension of aggregate d


backbone ( = 1.2)
Gp = Averaged elastic modulus of the aggregate with
respect to all its possible angular deformations (bend-
ing, twisting)
Gp is in fact controlled by the BdR which consists of a
layer of immobilized rubber, of thickness a
3 + FB
F: = 1.8 FB: = 1.2 :typical fractal dimensions for CB  = > = 3.5
3− F
Polymers and Carbon Black 219

A5.4.1  Mechanically Effective Solid Fraction of Aggregate


Np (D): = 200 : number of primary particle of size d in the aggregate of
size D
d: = 30⋅10 –9⋅m : primary particle diameter (e.g. = 30 nm in N330)
D: = 200⋅10 –9 ⋅m : aggregate size
a: = 2 ⋅ 10 –9⋅m : tightly BdR layer thickness (typical 2 nm)

3− F
 d
Aggregate solid fraction:    φ A (D) : =   φ A (D) = 0.103
 D

Solid volume of Np particles of  d3 


N p (D) ⋅  π ⋅  = 2.827 ⋅ 106  ( 10−9 ⋅ m )
3
diameter d:
 6

Diameter of equivalent sphere (according to Medalia) Des: =  d⋅Np(D)0.424

Des3
Volume of a sphere with diameter Des: π⋅ = 1.195 ⋅ 107  (10−9 ⋅ m)3
6

  d3  
 N p (D) ⋅  π ⋅ 6  
    = 0.237  = > the aggregate solid fraction as
 Des 3  ­calculated from fractal condiseration
 π ⋅ 6 
is likely underestimated

The effective fraction of the aggregate must include the tightly BdR fraction:

π 2⋅π 2   d  
N p (D) ⋅  ⋅ (d + 2 ⋅ a)3 − ⋅ a ⋅  3 ⋅  + a − a  
 6 3  2  
φe A (D) : =
π 3
⋅D
< =  Equation 71 given by
6 Heinrich and Klüppel
[Adv. Polym. Sci., 160, 1–44,
2002]
  d  
N p (D) ⋅ (d + 2 ⋅ a)3 − 4 ⋅ a 2 ⋅  3 ⋅  + a − a   N p (D) ⋅ (d 3 + 6 ⋅ d 2 ⋅ a + 6 ⋅ d ⋅ a 2 )
  2     or:   φe A (D) : =
φe A (D) : = D3
D3

 = > effective fraction of the aggregate : φe A(D) = 0.963

2⋅π 2   d  
The term ⋅ a ⋅  3 ⋅  + a − a  is expected to take into account the vol-
3  2  
ume resulting from the intersections of the rubber layer of thickness a, at the
220 Filled Polymers

contact points between two neighboring particles; however this corrective


term is in fact relatively small, as suggested by calculation results with the
typical values given above, i.e.:

π
⋅ (d + 2 ⋅ a)3 = 2.058 ⋅ 10 4  (10−9 ⋅ m)3 20, 580 nm 3
6

2⋅π 2   d  
⋅ a ⋅  3 ⋅  + a − a  = 410.501ϒ( 10−9 ⋅ m )
3
410 nm 3
3   2  

According to Heinrich and Klüppel an approximation for the effective aggre-


gate fraction is possible if a << d, as follows:

N p (D) ⋅ [(d + 2 ⋅ a)3 − 6 ⋅ d ⋅ a 2 ]


φefA (D) : = => φefA (D) = 0.965
D3

Rem: one notes that (d + 2⋅a)3–6⋅d⋅a2 = d3 + 6⋅d2 ⋅a + 6⋅d⋅a2 + 8⋅a3;


so this approximation means that the term 8.a3 is neglected

φefA (D) − φA (D) BdR increases the mechanically effective frac-


= 8.398 tion of the aggregate by some 8% (with respect
φ AD
to numerical values used)

φ = φA(D)   : space filling condition

Gp: = 1 ⋅ 1 09⋅Pa : elastic modulus of aggregate (para-crystalline


carbon: 1 GPa)

GR: = 1 ⋅ 106 ⋅ Pa : elastic modulus of a gum rubber (typical)

A5.4.2  Modulus as Function of Filler Volume Fraction

  3 + FB 
 φ: = 0, 0.01..0.25
  N p (D) ⋅ [(d + 2 ⋅ a)3 − 6 ⋅ d ⋅ a 2 ]   3 − F  
G(φ) : = GR + Gp ⋅  ⋅ φ  
 D3 
  Klüppel and Heinrich
model based on fractal
BdR : = 1, 1.01...1.4
considerations.

GGG(φ, BdR): = GR⋅[1 + 2.5⋅(BdR⋅φ) + 14.1⋅(BdR⋅φ)2] : Guth and Gold equation


Polymers and Carbon Black 221

10
Compound modulus, MPa

When compared with the


5 Guth and Gold equation
(which is based on mere
hydrodynamic consider-
ations) the Klüppel and
Heinrich model predicts
0
0 0.05 0.1 0.15 0.2 0.25 a stronger dependence on
Volume fraction filler volume fraction.
K&H G&G, 8% BdR G&G, 30% BdR

A5.4.3  Strain Dependence of Storage Modulus


[G. Heinrich, M. Klüppel. Adv. Polym. Sci., 160, 1–44, 2002. Equation 76]

−τ
  γ  2⋅m 
Ge( γ ) = (Ge0 − Ge f ) ⋅ 1 +    + Ge f
  γ c  

3 + FB
τ = elasticity exponent of percolat τ =
3− F
γ c = critical strain

Ge0 + Ge f
γ: = 0.0001, 0.0002..2   Ge0: = 8.4 ⋅ 106 ⋅ Pa   Gef: = 1.1 ⋅ 106 ⋅ Pa   = 4.75 ⋅ 106 Pa
2
Comparison with Kraus model

Heinrich and Klüppel


γcHK: = 0.20    mHK: = 0.22    τ: = 3.5
−τ
  γ  2⋅mHK 
GeHK ( γ ) : = (Ge0 − Ge f ) ⋅ 1 +    + Ge f
  γ cHK  
γ cK : = 0.007 m K : = 0.27
 1 
    Kraus : GeK (γ ): = ( Ge0 − Ge f ) ⋅  2⋅m K  + Ge f
1+  γ  
  γ cK  
222 Filled Polymers

10
γcK γcHK
Elastic modulus, MPa

The upper and lower limits of


5 the term
−τ
  γ  2⋅mHK 
1 +    are 1 and 0
  γ c  
0
1.10–4 1.10–3 0.01 0.1 1 10 respectively, and, at the critical
Strain strain, this term equals 0.088 ( = 2-τ)
Heinrich & Klüppel  = > the critical strain γcHK is the
H&K at critical strain
Kraus strain for which the elastic modu-
Kraus at critical strain lus is equal to Ge0⋅2-τ + Gef(1 – 2–τ).


1

−τ
  γ cHK  2⋅mHK 
1 +    = 0.088 –τ
  γ cHK  γ 2.mHK
 1 0.5
γcHK
2 −τ = 0.0088
2–τ

GeHK(γcHK) = 1.745⋅106 Pa
0
1.10–4 1.10–3 0.01 0.1 1 10
Ge0⋅2–τ + Gef(1 – 2–τ) = 1.745⋅106 Pa γ

A5.5  Lion et al. Model for Dynamic Strain Softening


[A. Lion, C. Kardelky, P. Haupt. Rubb. Chem. Technol., 76, 533–547, 2003]
A5.5.1  Fractional Linear Solid Model
(Rem: analog to the Zener model where the
dashpot viscosity has a ­fractional exponent)
Model parameters: ηβ

E1: = 6 ⋅ 106 ⋅ Pa : First spring modulus


E1
E2: = 31 ⋅ 106 ⋅ Pa : Second spring modulus
η: = 1.3 ⋅ 1010 ⋅ Pa⋅sec : fractional dashpot viscosity E2
β: = 0.495 : fractional exponent


Polymers and Carbon Black 223

η
λ:= λ = 419.355 sec : model characteristic timee
E2

ω : = 1 · Hz.. 100 · Hz  : frequency range  γ : = 0.0001, 0.0002…10  : strain range

Strain and frequency dependency of the intrinsic time scale:

b: = 51000   : constant   τ: = 1 · sec   : characteristic time

α: = 0.44    : fractional exponent

2
a( γ , ω ) : = 1 + ⋅ b ⋅ γ ⋅ (ω ⋅ τ)α
π

Intrinsic time scale vs. strain Intrinsic time scale vs. frequency
at fixed frequency at fixed strain

3.106 3.104
Intrinsic time scale
Intrinsic time scale

2.106 2.104

1.106 1.104

0 0
0 5 10 0 50 100
Strain Frequency
1 Hz 10 Hz 100 Hz Strain : 0.001 Strain : 0.01
Strain : 0.1

The fractional exponent a gives a nonlinear variation of the


intrinsic time scale with frequency (at constant strain).

A5.5.2  Modeling the Dynamic Strain Softening Effect


Elastic modulus:

2⋅β β
 λ   λ   π 
E2 ⋅  ω ⋅  +  ω ⋅  ⋅ cos  β ⋅  
 a( γ , ω )   a( γ , ω )   2  
Ge( γ , ω ) : = E1 + 2⋅β β
 λ   λ   π
1+ ω ⋅ + 2 ⋅ω ⋅ ⋅ cos  β ⋅ 
 a( γ , ω )   a( γ , ω )   2
224 Filled Polymers

π ⋅ (ω ⋅ λ − 1)
γ c (ω ) : =
2 ⋅ b ⋅ (ω ⋅ τ)α

4.107
γc(1) γc(100)
γ c (1 ⋅ Hz) = 0.013
Elastic modulus, Pa

3.107
Ge( γ c (1 ⋅ Hz), 1 ⋅ Hz) = 2.15 ⋅ 107 Pa
2.107
γ c (100 ⋅ Hz) = 0.17
1.107
Ge( γ c (100 ⋅ Hz), 100 ⋅ Hz) = 2.15 ⋅ 107 Pa
0
1.10–4 1.10–3 0.01 0.1 1 10
Strain
1 Hz 10 Hz 100 Hz 2 ⋅ E1 + E2
= 2.15 ⋅ 107 Pa
G' at critical strain 2

Viscous modulus:

β
 λ   π
E2 ⋅  ω ⋅ ⋅ sin  β ⋅ 
 a( γ , ω )   2
Gv( γ , ω ) : = 2⋅β β
 λ   λ   π
1+ ω ⋅ + 2 ⋅ω ⋅ ⋅ cos  β ⋅ 
 a( γ , ω )   a( γ , ω )   2

8.106
Gv( γ c (1 ⋅ Hz), 1 ⋅ Hz) = 6.349 ⋅ 106 Pa
γc(1) γc(100)
Viscous modulus, Pa

6.106
Gv( γ c (100 ⋅ Hz), 100 ⋅ Hz) = 6.349 ⋅ 106 Pa
4.106

2.106

0
1.10–4 1.10–3 0.01 0.1 1 10
Strain Frequency effects are well captured
1 Hz 10 Hz 100 Hz by the model.
Polymers and Carbon Black 225

Elastic modulus at 10 Hz

Elastic modulus at 10 Hz
4.107

γc(10)
3.107

2.107

1.107

0
1.10–4 1.10–3 0.01 0.1 1 10
γ
Ge(γ,10) Ge (γc(10.Hz), 10.Hz)
E1 E1+E2

According to the model, the upper and lower limits for the elastic modulus
are E1 + E2 and E1 respectively, and the mid modulus value corresponds to the
strain for the maximum value of the viscous modulus; the model exhibits
horizontal symmetry with respect to the mid modulus value.

Viscous modulus at 10 Hz

Viscous modulus at 10 Hz
8.106
γc(10)
6.106
Viscous modulus

4.106

2.106

0
1.10–4 1.10–3 0.01 0.1 1 10
Strain
G" at 10 Hz
max G" at the critical strain

The model gives a vertical symmetry of the G″(γ) curve with respect to the
critical strain for the maximum G″ value.
226 Filled Polymers

A5.5.3 A Few Mathematical Aspects of the Model


   π  π
cos  β ⋅  = 0.713 sin  β ⋅  = 0.702 : thesee two terms are close because of the particcular
 2  2
value of the fractional exponent β

For β = 0.5 they would be equal ( = 0.707)

2⋅β β
 λ   λ 
Terms  ω ⋅ and  ω ⋅ a( γ , ω )  vs. strain
 a( γ , ω ) 

Frequency = 10 Hz
1.103

γc(10)
100

10

0.1
At low strain, the 2 β term is the
dominant one;
0.01
at high strain, the β term is the
1.10–3 dominant one;
1.10–4 1.10–3 0.01 0.1 1 10 at the critical strain, both terms
Strain
2 beta Beta are equal.

2⋅β β
 λ   λ   π
Denominator 1 +  ω ⋅ + 2 ⋅ω ⋅ ⋅ cos  β ⋅  vs. strain
 a( γ , ω )  
 a( γ , ω )   2

1.104

1.103
Denominator

100

At high strain, the denominator


10 becomes asymptotic to 1.

1
1.10–4 1.10–3 0.01 0.1 1 10
Strain
at 10 Hz at 100 Hz
Polymers and Carbon Black 227

A5.6  Maier and Göritz Model for Dynamic Strain Softening


[P.G. Maier, D. Göritz. Molecular interpretation of the Payne effect. Kautsch.
Gummi, Kunstst., 49, 18–21, 1996]

A5.6.1  Developing the Model

Rubber–filler interaction sites 1 Local area on


are considered as “knots” carbon black particle
whose effect superimposes to Site occupied by
chemical networking. a stable "link"
The dynamic modulus is 2 Isolated site, available
proportional to the overall for an unstable "link"
networking density N and the Free site
temperature T (theory of rub-
ber elasticity)
3 4
Rubber chain
Ge = N⋅kB⋅T segments

kB = Boltzmann constant

Three contributions are considered for N: N = Nchem + Nstable + Nunstable

Nchem : vulcanization knots


Nstable : stable rubber-filler knots
Nunstable : unstable rubber-filler knots [ = f(strain)]

An equilibrium is assumed between adsorption and desorption of rubber


segments on filler sites, i.e.:

Φfree⋅vads = Φocc⋅vdes   [1]   Φfree: (volume) fraction of free sites

Φfree + Φocc = 1           Φocc: (volume) fraction of occupied sites

vads : adsorption rate, assumed to be constant


vdes : desorption rate, assumed to be proportional to strain amplitude,
i.e., vdes = K⋅γ [2]
Φ free v
It follows: Φ occ = ⋅ vads = (1 − Φ occ ) ⋅ ads
K⋅γ K⋅γ
1 1 [3]
=> Φ occ ( γ ) = =
K ⋅γ 1+ c⋅γ
1+
vads
228 Filled Polymers

The number of isolated sites Nisolated, available for unstable links, times the
fraction of free sites is in fact the number of unstable rubber-filler knots,
which depends on the strain amplitude, i.e.:

1
N unstable ( γ ) = N isolated ⋅ Φ occ ( γ ) = N isolated ⋅
1+ c⋅γ

There are thus, two contribution in the dynamic modulus, one due to the
vulcanization and the stable rubber-filler knots, the other owing to unstable
rubber-filler knots, thus, dependending on strain:

 1 
Gestable =  N chem + N stable + N isolated ⋅ ⋅kB⋅T
 1 + c ⋅ γ 
or:
1
Ge(γ ) = Gest + Geun ⋅
1+ c⋅γ

The viscous modulus is considered proportional to the fraction of occupied


sites times the probability P that an attached segment is able to slide into a
near (free) site, i.e. G″(γ)∼Φocc(g)P.
This probability depends on the quantity of free sites at the surface of the
particle, i.e., P∼Φfree(γ)

There are therefore, two contributions to the viscous modulus, one from the
stable links and one from the unstable links, the latter proportional to the
product Φfree × Φocc.

From Equations [1] and [2], one has K⋅γ 1 K ⋅γ


thus:   Φ free ( γ ) = Φ occ ⋅ = ⋅
vads K ⋅ γ vads
1+
vads

Therefore: Gv(γ) = Gvstable + Gvunstable⋅Φocc(γ)⋅Φfree(γ)

1  1 K ⋅γ 
Gv(γ ) = Gv stable + Gv unstable ⋅  ⋅
K ⋅γ K ⋅ γ vads 
1+ 1+ 
vads  vads 

c⋅γ K
or:  Gv(γ ) = Gv stable + Gv unstable ⋅ with c=
(1 + c ⋅ γ )2 vads
Polymers and Carbon Black 229

A5.6.2 A Few Mathematical Aspects of the Model


γ: = 0.0001, 0.0002, .. 2 : strain range for calculation

Gest: = 0.74 ⋅ 106 ⋅ Pa   c: = 40.15   Gvst: = 0.10 ⋅ 106 ⋅ Pa Data for butyl/N330 cpd
(f N330 = 0.233) used by Maier
and Göritz when probing
their model
Geun: = 9.52 ⋅ 106 ⋅ Pa   Geun: = 4.55 ⋅ 106 ⋅ Pa

1 c⋅γ
Ge(γ ): = Gest + Geun ⋅ Gv(γ ):= Gv st + Gv un ⋅
1+ c⋅γ ( 1 + c ⋅ γ )2
1.5.107 1.5.106
1 1
c
c
Viscous modulus, MPa
Elastic modulus, MPa

1.107 1.106

5.106 5.105

0 0
1.10–4 1.10–3 0.01 0.1 1 10
1.10–4 1.10–3 0.01 0.1 1 10
Strain Strain

c is the reverse of the strain for c is the reverse of the strain for
which G′ = 0.5 × G′unstable + G′stable. which G″ is maximum and equal
to 0.25 × G″unstable + G″stable.

 1  1
Ge   = 5.5 ⋅ 106 Pa Gv   = 1.238 ⋅ 106 Pa
 c  c

Gest + 0.5 ⋅ Geun = 5.5 ⋅ 106 Pa Gv st + 0.25 ⋅ Gv un = 1.2238 ⋅ 106 Pa

1 −c
Rem: one notes also that the 1st derivative of is
1+ c⋅γ ( 1 + c ⋅ γ )2
Due to its starting hypotheses, the Maier and Göritz model has a mathe-
matical form that leads to symmetries in both the G′ vs. strain and the G″ vs.
strain functions; the former exhibits an horizontal symmetry with respect to
a mid- modulus value, the latter shows vertical symmetry with respect to a
critical strain = 1/c.
Both the elastic and the viscous moduli at the critical strain are simple
combinations of the “stable” and “unstable” links contributions.
230 Filled Polymers

A5.6.3  Fitting the Model to Experimental Data


Dynamic strain softening data on SBR/60 phr black cpds
[M. Gerspacher, C.P. O’Farrell, C. Tricot, L. Nikiel, H.A. Yang. ACS Rubb. Div;
Mtg. Louisiana, KY, 1996. Paper 74]

Strain sweep experiments Maier and Göritz model


Elastic modulus
N660 N330 N110 1
Ge( γ ) = Gest + Geun ⋅
Strain G′ G ′′ G′ G ′′ G′ G ′′ 1+ c⋅γ
% MPa MPa MPa MPa MPa MPa

 0.091 2.34 0.265 4.45 0.630 6.02 0.915  Gest = G′stable: elastic modu-
  lus due to chemical cross-
 0.19 2.31 0.275 4.41 0.630 5.95 0.938 
 0.28 2.330 0.280 4.30 0.650 5.80 0.942  links + stable rubber-filler
 
 0.38 2.26 2.290 4.24 0.660 5.61 0.960  interactions
 0.48 2.24 0.295 4.10 0.670 5.50 0.9980 
  Geun = G′unstable: elastic
 0.55 2.23 0.296 4.05 0.685 5.32 0.993 
  modulus due to unstable
 0.66 2.22 0.297 3.95 0.692 5.18 1.000 
 0.74 2.21 0.297 3.89 0.7700 5.05 1.010  rubber–filler interactions
 
 0.84 2.20 0.297 3.82 0.710 4.95 1.026  [ = f(strain)]
 0.93 2.18 0.298 3.75 0.716 4.83 1.030 
 
 1.07 2.16 0.2298 3.61 0.720 4.65 1.032  Viscous modulus
 
Data: =  1.22 2.15 0.299 3.55 0.720 4.48 1.033  c⋅γ
Gv( γ ) = Gv st + Gv un ⋅
 1.48 2.14 0.299 3.46 0.719 4.31 1.031  ( 1 + c ⋅ γ )2
 
 1.7 2.12 0.298 3.40 0.718 4.20 1.030 
 1.9 2.11 0.298 3..34 0.717 4.08 1.020  Gvst = G″stable: viscous mod-
 
 2.8 2.06 0.294 3.08 0.690 3.71 0.968  ulus due to chemical cross-
  links + stable rubber-filler
 3.7 2.05 0.290 2.95 0.660 3.49 0.914 
 4.6 2.04 0.294 2.85 0.631 3.30 0.870  interactions
 
 5.5 2.02 0.285 2.78 0.608 3.18 0.828 
Gvun = G″unstable: viscous
 6.5
 2.00 0.280 2.71 0.581 3.08 0.782 
 7.6 1.99 0.275 2.65 0.560 3.01 0.752 
modulus due to unstable
  rubber–filler interactions
 8.5 1.97 0.271 2..62 0.542 2.95 0.722 
 9.5 1.95 0.269 2.57 0.530 2.90 0.700  [ = f(strain)]
c = constant related to
the rate of adsorption-
desorption of rubber seg-
ments on appropriate sites
on filler particles
γ = strain amplitude
Polymers and Carbon Black 231

A5.6.3.1 Modeling G′ vs. Strain


Extracting data Strain(%) N660G ′(MPa) N330G′(MPa) N110 G′(MPa)
  j: = 0..2 γ : = Data <0> Ge0 := Data <1> Ge1 := Data <3> Ge2 := Data<5>

n: = length( γ ) n = 23 : number of data

Guess parameters for nonlinear fitting algorithm (GenFit function)

 ( Ge j )0 
< = extracting guess parameters from experi-
  mental data, i.e.,
C j : =  ( Ge j )n− 1  - The highest measured modulus for G′stable
  - The lowest measured modulus for G′unstable
 1 
γ  - The reverse of the mid range strain for c
 n
 round 2    C1 
Model equation and partial  C0 + 1 + C ⋅ γ 
 2

derivatives:  1 
 
F ( γ , C) : =  1 
 1 + C2 ⋅ γ 
 
 −C1
⋅γ
 ( 1 + C ⋅ γ )2 
 2 

      N660 N330 N110

 2.34   4.45   6.02 


     
C0 =  1.95      C1 =  2.57      C2 =  2.9 
 0.676   0.676   0.676 

Resj: = GenFit(γ,Gej,Cj,F) : calling the nonlinear regression algorithm

Fitting equation Correlation coefficient r2


 

Ga j : = ( Re s j )0 +
( Re s j )1  
Gar2 j : = corr  ( Res j )0 +
( Re s j )1 , Ge 
 
1+ ( Res j )2 ⋅ γ  1 + ( Re s j )2 ⋅ γ  
j
  
 

   N660 N330 N110


:Gstable
1.917   2.236   2.328 
Res 0 =  0.445     Res1 =  2.421     Res 2 =  4.072     :Gunstable

 0.708   0.642   0.679  :c

   Gar20 = 0.997 Gar21 = 0.999    Gar22 = 0.999 :r 2


232 Filled Polymers

Elastic modulus, MPa


4

0
0.01 0.1 1 10 100
Strain, %

A5.6.3.2  Modeling G″ vs. Strain

Extractingdata Strain(%) N660 G ′′(MPa) N330 G ′′ ( MPa) N110G ′′ ( Mpa)



j = 0..2 γ : = Data<0> Gv 0 := Data<2> Gv1 := Data<4> Gv 2 := Data<6>

Guess parameters for nonlinear fitting algorithm (GenFit function)


Finding max G″ and corresponding γ

gc j : = A ← max ( Gv j ) Extracting guess parameters  


 Gv
for i ∈1..n − 1
from experimental data, i.e.: ( j )n−1 
- The measured modulus at  
C j : ( gc j )1 
 γi  the highest strain for G″stable
 if ( Gv j )i = A
B←  
( Gv j )i  - The highest measured  1 
 
A←B
modulus for G″unstable  ( gc j )0 
- For c, the reverse of the
A strain corresponding to the
max G″ data

 C ⋅C ⋅ γ 
 c0 + 1 2 2 
 (1 + C2 ⋅ γ ) 
 1 
Model equation and  
partial derivatives: Q( γ , C) : =  C2 ⋅ γ 
 
 (1 + C2 ⋅ γ ) 2

 C1 ⋅ γ γ2 
 − 2 ⋅ C1 ⋅ C2 ⋅ 3
 ( 1 + C2 ⋅ γ ) (1 + C2 ⋅ γ ) 
2
Polymers and Carbon Black 233

       N660 N330 N110

 0.269   0.53   0.7 



C0 =  0.299   
C1 =  0.72   C2 = 1.033 
 0.676   0.82   0.82 
       

Resj: = GenFit(γ,Gvj,Cj,Q) : calling the nonlinear regression algorithm

Fitting equation Correlation coefficient r2


 


Gb j : = ( Res j )0 +
( Res j )1 ⋅ ( Res j )2 ⋅ γ  
Gbr 2 j : = corr  ( Res j )0 +
( Res j )1 ⋅ ( Res j )2 ⋅ γ , Gv 
j
 1 + ( Res j ) ⋅ γ    1 + ( Res j ) ⋅ γ  
2 2
 2   2 

   N660 N330 N110

 0.254   0.492   0.661  : G′′ stable


Res 0 =  0.188  Res1 =  0.903  Res 2 = 1.507  : G′′ unstable
 0.799  1.027   1.243  :c
        
   Gbr20 = 0.968 Gbr21 = 0.925 Gbr22 = 0.92 : r2

1
Viscous modulus, MPa

The Maier and Göritz model fits


0.5 reasonnably well the experimental
G′ data but cannot meet the asym-
metric shape of G″ vs. strain data.

0
0.01 0.1 1 10 100
Strain, %
6
Polymers and White Fillers

6.1  Elastomers and White Fillers


6.1.1 Elastomers and Silica
6.1.1.1  Generalities
In contrast with carbon black, silica has particular surface properties that
bring a number of problems when using such materials as reinforcing fillers,
particularly in hydrocarbon elastomers. Indeed the surface of silica, either
fumed or precipitated, is strongly polar and hydrophilic, owing to its polysi-
loxane structure with numerous silanol groups. This particular surface
chemistry of silica has several immediate consequences:

1. Silica surface can adsorb significant quantities of water, as reflected


by the well known usage of the material as drying agent (or mois-
ture absorber)
2. Moist silica is very difficult to dry
3. Interparticle interactions are very strong because of hydrogen
bonding
4. Silica is the ideal filler for silicone polymers, for instance
polydimethylsiloxane
5. When used as filler for diene elastomers, chemical modification
of silica surface is required firstly to promote mixing through
decreased inter-particles interactions and secondly to establish
­adequate ­rubber–filler interactions.

Most of the specific properties of silica (and silicates) were known more than
three decades ago, as reviewed by Wagner in 1976,1 who somewhat foresaw
the tremendous developments that followed. Indeed, in the last decades, sig-
nificant progress was made in using silica in diene elastomers, supported by
the peculiar dynamic properties this filler brings to (diene) elastomers. It per-
mitted the development of the so-called “green tire,” because a lower rolling
resistance is obtained, when compared with carbon black reinforced tires.

235
236 Filled Polymers

6.1.1.2  Surface Chemistry of Silica


The siloxane group Si–O–Si is nonpolar and therefore would provide a hydro-
phobic character to the silica surface. But, as illustrated in Figure 6.1, there
are various types of hydroxyl groups that are strongly hydrophilic, from iso-
lated (or free) silanols to silanediols and silanetriols (rare, if any). All these
groups are easily identified by infrared analysis. Vicinal silanols can develop
hydrogen bonding and eventually siloxane groups through water elimina-
tion. Free silanols are more reactive than vicinal groups and therefore pro-
mote inter-particles hydrogen bonding. They are also the prime reactive site
for organic molecules, namely organo-silanes. The average surface density is
around 3–6 silanol/nm2, depending on the specific area. Thanks to various
equilibrium reactions involving hydrogen interactions, silica can adsorb sig-
nificant quantities of water, in such a manner that all silanol groups on the
surface are easily saturated. When silica is heated above 100°C, such a phys-
isorbed water is removed and becomes negligible above 250°C. Between 200
and 500°C, condensation of vicinal silanols occurs to yield siloxane groups,
and above 600°C, free silanols start also to condensate.
When compared with carbon black, the surface chemistry of silica is the key
parameter in reinforcement with this filler. Indeed silica surface is occupied
by sizable quantities of siloxane and silanol groups, giving rise to hydrogen
interaction with either “free” or “bound” water. Free water is easily removed
by drying at 105–250°C, while bound water is only released at 900–1000°C
and results in fact from the condensation of vicinal silanol groups. As can be
expected the free moisture content of silica strongly affects the rheological
and curing properties of silica filled rubber compounds. Another important
consequence of the oxygen rich surface chemistry of silica are the strong
interparticles interactions through hydrogen bonding that, on one hand
give rise to poorer dispersibility than carbon black (in nonpolar polymers),
and on the other hand permit the use of suitable chemical promoters, e.g.,
bis(triethoxysilylpropyl)tetrasulfane, TESPT, to form covalent bonding with
(unsaturated) elastomer. Of course, certain specialty elastomers such as

H H
OH
O O OH OH OH OH OH O

Si Si Si Si Si Si Si Si Si

O O O O O O
O O O O O O O O O O O O O O
Vicinal silanols Free silanol Silanediol Siloxane
=> hydrogen bonding Silanetriol
between particles (rare, if any)

Figure 6.1
Surface chemistry of silica.
Polymers and White Fillers 237

polydimethylsiloxanes naturally interact with silica, thank to their similar


chemistry. It is for instance long known that, with silicone polymers, the silica
surface chemistry can be varied nearly at will, by controlling the degree of
adsorbed water, the hydroxyl population and the degree of organophilicity,2
all aspects largely exploited by the silicone rubber industry.3

6.1.1.3  Comparing Carbon Black and (Untreated) Silica in Diene Elastomers


Silicone rubber and, in general polar polymers, are by nature materials of
choice for preparing silica filled systems; however limited to niche applica-
tions, with respect to the range of properties that such specialty polymers
may offer. In order to develop optimum reinforcing performance with more
common diene elastomers, silica must be chemically treated as we will see
below, because contrary to carbon blacks, silica particles do not develop
spontaneous strong interactions with nonpolar polymers. It is nevertheless
interesting to see that, even with comparable size and structure, pure silica
does not affect the mechanical properties of vulcanized rubber compounds
in the same manner as carbon black.
This was clearly demonstrated in the excellent review paper by S. Wolff 4
who studied the effects of two comparable series of silica and carbon black
in 50 phr filled natural rubber (RSS1) compounds, vulcanized with perox-
ide (Note that such a vulcanization system was chosen because there is no
interference between silica and peroxide curing). Table 6.1 gives typical size
and structure data for the two series of filler considered. It is worth under-
lining that there are no standard methods for characterizing silica. Either an
existing method can be used as such because it does not depend on the filler

Table 6.1
Comparable Series of Precipitated Silica and Carbon Black
DBP or TEA Absorptiona
N2
(ml/100 g)
Specific Area
Filler Grade (m2/g) Uncompressed Crushed (24M4)
Precipitated silica 1 30 100 64
2 48 164 74
3 123 192 90
4 167 227 90
5 172 188 96
6 173 204 93
Furnace blacks N660 36 95 70
N550 40 123 86
N326 76 70 64
N330 78 100 85
N356 88 153 113
N220 110 114 94
N110 139 115 94
a DBP, di-butylphthalate for carbon black; TEA, triethanolamine for silica.
238 Filled Polymers

considered, or the method must be modified to take into account the surface
chemistry of silica. For instance the specific area of silica can be assessed
through nitrogen adsorption (BET method) but for aggregate structure, di-
butylphtalate adsorption is not convenient (because DPB does not “break”
interparticles hydrogen bonding). Adsorption of triethanolamine gives cor-
rect results, comparable to data obtained on carbon black with DPB.
Comparing mechanical properties imparted by either carbon black or
silica in a purposely simple natural rubber formulation allows several inter-
esting conclusions to be drawn. Figure 6.2 shows for instance the 100% and
200% modulus, both affected by the size and the structure of the filler. At low
strain, i.e., 100%, most precipitated silica and several high structure blacks
exhibit similar reinforcing capabilities; at higher strain however, all silica
are clearly less reinforcing than carbon black. Wolff attributed this effect to a
“silica network” which is destroyed when straining vulcanizates.
Low strain amplitude dynamic properties reveal quite an interesting
aspect of silica reinforcement. As shown in Figure 6.3, high structure silica
(i.e., with crushed TEA adsorption values higher than 80 ml/100 g) give NR
compounds with higher complex modulus G* and lower tan δ than carbon
blacks of similar structure. This advantages of high structure silica over car-
bon blacks is also observed when performing technological dynamic tests,
for instance rebound resilience test.
However, in line with tensile modulus data, silica gives larger compres-
sion sets than carbon blacks. Despite the fact that the compounds investi-
gated were (purposely) oversimplified with respect to industrial practices,
the key information in the experiments reported by Wolff is that silica filled
compounds exhibit definitely better dynamic properties that carbon black
filled ones, namely higher rebounds and lower heat build-up. In addition
the higher the specific area of fillers, the larger the differences between silica
and carbon black loaded materials. Freund and Niedermeier made a similar

4.5 20 NR (RSS1) 100


Carbon black Filler 50
4.0 Silica Peroxide (DCP) 2.03
200% Modulus, MPa
100% Modulus, MPa

15
3.5
3.0
10
2.5
2.0 5
1.5
1.0 0
40 60 80 100 120 40 60 80 100 120
Crushed DBP or TEA absorption, ml/100g Crushed DBP or TEA absorption, ml/100g

Figure 6.2
Effect of carbon black and silica structure of tensile properties. (Drawn using data from S.
Wolff, Rubb. Chem. Technol., 69, 325–346, 1996.)
Polymers and White Fillers 239

32 0.20 NR (RSS1) 100


Carbon black Filler 50
Complex modulus E*, MPa

Silica Peroxide (DCP) 2.03


26
0.15
Frequency : 5 Hz

Tan delta
20
Temperature : 23°C
0.10
14
0.05
8

2 0
40 60 80 100 120 40 60 80 100 120
Crushed DBP or TEA absorption, ml/100g Crushed DBP or TEA absorption, ml/100g

Figure 6.3
Effect of carbon black and silica structure of low strain dynamic properties. (Drawn using data
from S. Wolff, Rubb. Chem. Technol., 69, 325–346, 1996.)

comparative study on carbon black and silica compounded in nonpolar and


polar elastomers.5 From their investigation of the dynamic strain softening
effects, they concluded that reinforcement by carbon black and silica essen-
tially proceed from different micro-mechanisms. Polymer adsorption pre-
vails in carbon black filled systems, while filler particle networking is the
key aspect in silica filled systems.
Performed some 20 years ago, such basic studies (and many others) clearly
indicated that highly structured silica (i.e., very large specific area) were
surely interesting alternative fillers for carbon black in highly demanding
dynamic applications, for instance tire technology. However, when compared
to carbon blacks, silica have serious drawbacks, essentially arising from their
peculiar surface chemistry. Besides the strong inter-particles interactions
which give dispersion difficulties in hydrophobic polymers (notably all diene
rubbers, i.e., around 90% of the overall elastomers consumption), the chemi-
cally active surface of silica has a strong potential for interacting/interfering
with curing systems, particularly when basic accelerators are used, giving
lower cure rates and lower crosslink densities (i.e., modulus).4,6,7

6.1.1.4  Silanisation of Silica and Reinforcement of Diene Elastomers


In order to fully exploit the promising capabilities of silica in the reinforce-
ment of diene elastomers, it is essential to consider their surface chemistry
and to accordingly proceed to a number of changes, at various levels of rub-
ber technology, in terms of formulation, compounding, mixing, and pro-
cessing. First the vulcanization chemistry must be modified in order to take
into account the high chemical reactivity of silica particles surface and the
likely modification of the structure of the networking bonds, with respect to
the experience gained over the years with carbon blacks. Second particle–
particle interactions are very strong with silica, due to hydrogen bonding
240 Filled Polymers

and the hydrophobic character of diene elastomers is surely not a favorable


aspect in what silica dispersion is concerned. Third rubber–filler interactions
of physical origin, as occurring with carbon blacks, cannot be expected with
silica with respect to the shielding effect of silanol groups. It follows that a
chemical approach has to be considered in order to create covalent bonding
between rubber and silica particles.
The benefit in using silanes as coagents in (diene) rubber compounding
has been recognized for long1,8 and many organo-silanes were studied. A
number of investigated organo-silanes were found either limited in their
interest, difficult or inconvenient to use, or too expensive for industrial
applications. Eventually, the reduction of silica interparticle interactions, in
association with the development of suitable (chemical) bonds with diene
elastomers were obtained through the use of so-called “reinforcement pro-
moters,” essentially bifunctional silanes of general formula:

(RO)3–Si–(CH2)n–X

One end of such molecules is expected to specifically react with silanols on


silica surface, whilst the other end is expected to eventually interact with the
vulcanization system (essentially sulfur based) in order to provide chemical
bonding with the rubber network. It worth underlining again the fundamen-
tal difference between carbon black and silica reinforcement: no chemistry is
needed with the former but is essential with the latter.
Many organo-silanes have been synthesized and tested over the years
(mainly in the 1970s) and essentially two chemicals were found of interest in
the rubber (tire) industry:

• Bis(3-triethoxysilylpropyl)tetrasulfane (TESPT); note that TESPT is


in fact a mixture of different polysulfane with an average S chains
of four9)

(C2H5O)3–Si–(CH2)3–S4–(CH2)3–Si–(OH5C2)3

frequently referred under its commercial name Degussa (Evonik)


Si69
• 3-thiocyanatopropyl-triethoxy silane (TCPTS)

(C2H5O)3–Si–(CH2)3–SCN

also referred under the commercial name Degussa (Evonik) Si264.

In principle, one may either pretreat silica with such silanes (usually in
solution/suspension, with subsequent elimination of solvents) and then use
the modified silica in compounding, or consider silane as a compounding
Polymers and White Fillers 241

ingredient and proceed to the silanization during mixing operations.


Obviously pretreated silica are expensive products since notwithstanding
the cost of such fine chemicals as organo-silanes, solvent elimination, prod-
uct drying and conditioning bring uncompressible costs. In situ silanization
became therefore the preferred approach despite the challenging difficulties
in controlling a chemical reaction in a highly viscous medium, i.e., during
mixing. In other terms, quite a complex set of physico-rheological processes
had to be mastered in equipment that at first were essentially developed for
preparing carbon black compounds. Chemistry in highly viscous media is
not an issue in carbon black reinforcement and therefore, controlling the
variation of temperature during mixing is essentially considered in terms
of limitation of the warming up associated with the process, mainly due to
viscous heat dissipation effects as arising when shearing a viscoelastic mate-
rial. With in situ silanization of silica (i.e., during mixing), the problem is
completely different since elementary considerations allows the following
requirements and difficulties to be a priori identified:

1. An even dispersion of all reactive ingredients has first to be achieved


in a highly viscous medium, which means than reactive species dis-
placement is an issue.
2. Reactions between silica and silane must be activated (usually by
reaching the appropriate temperature) and completed (by maintain-
ing the appropriate temperature conditions for a sufficient time.
3. Premature reactions between the rubber and the silane must be
avoided during mixing operation.

Directly studying silanization chemistry during rubber mixing is very


challenging and has never been made (or even tempted) to the author’s
knowledge. But a number of very elegant studies of silanization in solution
or in suspension have been performed9–13 that eventually confirmed earlier
proposals for a likely reactional scheme,14 and allowed to understand cer-
tain aspects of the in situ process and the interference with vulcanization.
Investigations using rubber compounds15 essentially confirmed the conclu-
sions of such basic studies.
The silica modification with a bi-functional organosilane (either TESPT of
TCPTS for instance) and the subsequent development of rubber–filler bond-
ing during vulcanization is essentially considered as follows:16

1. Silanisation (Figure 6.4): first one ethoxy group reacts quickly with
an isolated silanol (around 85% on silica surface) or a silanediol (15%);
then there is hydrolysis of the remaining ethoxy groups, which pro-
duce a reticulation of silane molecules through siloxane bonding.
2. Vulcanization (Figure 6.5): the tetrasulfane group (with TESPT silan-
ated silica) is broken and forms rubber–filler covalent bonds with the
242 Filled Polymers

Si O OC2H5
Si
OH C2H5O OC2H5 -CHOH O
Si Si (CH2)3 Sa
+ C2H5O Si (CH2)3 S4 (CH2)3 Si OC2H5
Si OH C2H5O Si O (CH2)3 Sb
OC2H5 Si
Si O OC2H5
Silica reaction with TESPT Silanated silica
(a+b = 4)

Figure 6.4
Silica modification with bi-functional organosilane.

Si O OC2H5 Si O OC2H5
Si Si
Si O (CH2)3 Sa Sulphur S8 Si O (CH2)3 Sa
+ Rubber
Si O (CH2)3 Sb Accelerator Si O (CH2)3 Sb
Si Si
Si O OC2H5 Si O OC2H5

Figure 6.5
Silica–rubber bonding during vulcanization.

polymer during the rubber networking. Note that bonds between


silanated silica and rubber are either mono- or disulphidic.

In fact, the silanization itself occurs in two steps: first there is a reaction
between the silanol groups on silica surface with the alkoxy group of the
silane, likely through hydrolysis of the alkoxy groups followed by a con-
densation reaction with the silanols, but direct condensation is also possible.
Hydrolysis then condensation is supported by the beneficial influence of the
moisture content of silica on the rate of silanization. The second step is a
condensation reaction between adjacent molecules of the silane on the silica
surface, and a hydrolysis step is also likely occurring. The result is a sig-
nificant decrease of the hydrophilic degree of silica particles and hence an
easier dispersing in hydrophobic elastomers (i.e., most diene rubbers used
in tire technology). Detailed investigations on the kinetics of this complex
set of reactions have demonstrated that the activation energy of the first
step is nearly twice of what is needed for the second step (i.e., 47 kJ/mole vs.
28 kJ/mole) but the secondary step is around 10 times slower that the first
one.13 Recently reported results, obtained by using a model silane and time
resolved IR spectroscopy in a microreactor with infrared transparent win-
dows brought a very elegant confirmation of such a two steps mechanism.17
It was indeed shown that the silane interacts first by hydrogen bonding with
isolated silanol groups. This first step is very fast and the so-immobilized
species react dissociatively with silanols to give covalent bonding with the
silica surface, while alcohol is released. Hydrogen bonded silane is less stable
Polymers and White Fillers 243

than when covalent bonded. It was also found that vicinal silanol groups do
not react with silane, likely owing to a lower reactivity or to steric hindrance.
Such results explain why only 25% of the total hydroxyl groups on silica sur-
face are involved in the silanization process. Both sequences of reactions are
acid as well as alkaline catalyzed and the rate constant for the primary reac-
tion decreases as the silane content increases. It is likely that the lower acces-
sibility of silanol groups on the filler surface as the silane content increases,
and the decrease of H2O available for the hydrolysis step are responsible for
this effect. It follows that the optimal loading of the silane is around eight
parts of TESPT for 100 parts of silica. Details on this complex chemistry can
be found in the referenced papers.
The reactional scheme described above prompts several remarks: first eth-
anol is produced during silanization reactions (around 2 moles of ethanol
per mole of TESPT), and must be eliminated of captured by the appropriate
formulation ingredient, otherwise there will be porosity in the vulcanized
product; second with TESPT, highly reactive tetrasulfane groups are formed
which may give thermo-activated reactions with the polymer if the tempera-
ture is too high. It follows that controlling the temperature during mixing is
a crucial aspect of the operation: it must be high enough for the silanization
reaction to be activated and low enough for tetrasulfane networking or pre-
mature vulcanization to be avoided. As may be expected the nature of the
alkoxy group in the organo-silane is playing a role in the silanization process
(Figure 6.6); whilst very fast, the methoxy group cannot be used for obvious
toxicological reasons and the reaction rate is decreasing with the size of the
group, leaving the ethoxy as the best choice. Ethanol formation during the
in situ silanization is readily an issue in practical compounding since for
each gram of silane used, around 0.5 g of alcohol would be produced if all
ethoxy groups were reacting. Not all ethoxy groups are reacting however13
but, on the factory floor, considerable amount of ethanol are produced, which
besides potential health and toxicity hazards, can readily decrease the effi-
ciency of the mixing process, through recondensation in the mixer chamber
and hence wall slippage of the compound. The nature of the rubber has been
found to affect the silica silanization process, as demonstrated by Table 6.2.
It was also established that, at constant mixing time, the reaction efficiency
increases with the (dump) temperature, but in the mean time higher mix-
ing temperature increases the risk of premature vulcanization. Lengthening
the mixing cycle and/or using several (re)mixing steps offers several advan-
tages, most likely because it favors the volatilization of ethanol. Dierkes has

CH3O- > C2H5O- > C3H70- > C4H9O > ...


Methoxy group not used
for toxicological reasons Decreasing rate

Figure 6.6
Effect of the alkoxy groups in the silanization efficiency of organo-silanes.
244 Filled Polymers

Table 6.2
Effect of Rubber Type on the Silica Silanisation Reaction
Rubber or Rubbers Blend Mole Ethanol/mole TESPT
S-SBR/BR 1.50
NR 1.75
NR/BR 1.80
E-SBR 2.25
NBR 0.90
EPDM 2.30
Source: Data from U.Görl and A.Parkhouse Kautch. Gummi
Kunstst., 52, 493–500, 1999.
Note: Experimental conditions: TESPT content: 6.5% of silica; 5
min. mixing with dump at 160°C.

40 0.30
Frequency : 5 Hz Frequency : 5 Hz
Temperature : 23°C Temperature : 23°C
Natural rubber cpd N110 carbon black
30 Silica
Elastic modulus E´, MPa

0.20
Tan δ

20

0.10 Silica + TESPT


N110 carbon black
10

Silica
Silica + TESPT
0 0
–3 –2 –1 0 –3 –2 –1 0
Log double strain amplitude Log double strain amplitude

Figure 6.7
Effect of silanization on the reinforcing properties of NR compounds (Data from S. Wolff,
Rubb. Chem. Technol., 69, 325–346, 1996.)

thoroughly investigated the industrial mixing of silica filled compounds and


considered several approaches to overcome such practical problems.18
Silanisation has profound effects on the reinforcing character of silica
and allows to obtain vulcanized rubber systems which exhibit certain
benefits with respect to corresponding carbon black filled compounds. As
expected silica interparticle interactions are considerably reduced through
silanization, as reflected by the large reduction in the dynamic strain soft-
ening effect. Figure 6.7 shows for instance the dynamic properties of NR
Polymers and White Fillers 245

compounds filled with carbon black, silica, or silica + TESPT, through strain


sweep experiments performed at 5 Hz frequency.
As can be seen, the treatment of silica with TESPT reduces the elastic
modulus drop upon increasing strain amplitude; this indeed corresponds
to a strong reduction of interparticles interactions but the reinforcing effect
is also reduced (lower modulus that the reference carbon black filled com-
pound. However the interest in using treated silica appears on tan δ which
is slightly increased through silanization but remains significantly lower
than the homologous carbon black filled system. In other words, silanization
reduces the dynamic strain softening effect but keeps the lower viscous dis-
sipation of silica under dynamic strain. These characteristics of silane treated
silica are at the origin of the development of silica filled tread bands in auto-
motive tires (so called “green tires” because with the lower viscous dissipa-
tion imparted by silica reinforcement, the energy produced by the engine is
more efficiently used in moving the vehicle).
The complex chemistry associated with the in situ silanization of silica, as
mastered today by tire manufacturers, is surely worth detailed considerations,
particularly with respect to the processing behavior. The mixing procedure
and conditions, the temperature control, the elimination of ethanol, etc. are key
issues on the factory floor, so-far typical of each mixing plant and essentially
monitored through pragmatic engineering practices, about which tire manu-
facturers remain relatively discreet. Silanols can react with several compound-
ing ingredients such as stearic acid, polyalcohols, and amines and, obviously
would compete with silane and reduce the silanization efficiency. The order
of addition of compounding ingredients is consequently of prime importance;
indeed the mixing procedure must be such that the whole amount of TESPT is
consumed during the primary reaction.15 In situ silanization is a very important
subject in contemporary rubber technology, actually well mastered through
the appropriate (and complicated) engineering practices,7,19 but outside the very
scope of this book. What must be kept in mind with respect to the scope of
this book, is first that for silanes to be effective “promoters” of mineral fillers,
active chemical groups on filler particles are needed in quite large quantities,
second that specific conditions must be met and maintained for a sufficient
time, for the silanization to be complete, third that the polymer must have a
reactivity potential with some chemical functionalities of the silane. Silica is
surely the right filler for the effective use of silanes, but using such chemicals
with other minerals having no or a poorer surface chemistry (e.g., kaolin, mica,
talc, …) is obviously not expected to bring the same benefits on reinforcement.
Nevertheless, as coated mineral particles are generally easier to dispersed in
diene rubbers than their uncoated equivalent, silane-treated talc, mica, and
kaolin grades permit fine dispersion to be obtained, with obviously less clus-
tered particles that are always potential failure initiation sites.
Any silica-filled rubber formulation may benefit from in situ silanization
during the compounding operations and, in the tire industry, it is now well
established that partial or total substitution of carbon black by silanated
246 Filled Polymers

(precipitated) silica in tread formulation gives the best compromise in terms


of rolling resistance and wet grip. It is interesting to note that most of the
­silica-filled tread compounds in use today are essentially developments of
the original Michelin formulation patented by Rauline,20 in which a medium-
or high-vinyl solution styrene butadiene copolymer (S-SBR) is the main
elastomer. S-SBR is produced by anionic polymerization whose capabilities
to control the molecular weight distribution and the level of branching are
well known. All such characteristics are of course of importance as they pro-
vide the rubber matrix the required performances with respect to tire tread
dynamic behavior. However, as recently pointed out by Heinrich and Vilgis,21
the reason why S-SBR (instead of less expensive emulsion E-SBR) is the poly-
mer of choice for silica-filled tread compounds is not fully understood. When
comparing the tan δ vs. dynamic strain amplitude functions (at 30°C) of two
S-SBR and E-SBR compounds with equal silica content (80 phr) and the same
hardness, these authors note that the S-SBR system has a significantly lower
tan δ (around 20% drop). Such a difference can hardly be explained with
respect to the microstructure (cis, trans and 1,2-vinyl) and MW distribution
of both elastomers, and Heinrich and Vilgis develop an argument based on
the confinement of polymer segments in pores that would exist in amorphous
precipitated silica, seen as clusters of elementary spherical particles. Polymer
segments could be immobilized (or confined) in such pores if, indeed, their
size, in the nanometer range, is close to typical dimensions of the polymer
chain, for instance the so-called Kuhn length (from the theoretical view of
a polymer chain seen as made of N segments of Kuhn length b, so that each
segments are freely jointed with each other; the so-called contour length of
the polymer is then L = N b). Because E-SBR is somewhat more branched than
S-SBR, it would saturate the external surface of silica clusters, without pen-
etrating much into silica pores. On the reverse, S-SBR segments would have
the capability to completely fill pore volume and remained confined, thus giv-
ing strong polymer–filler interaction, that would of course superimposed to
the chemical interaction imparted by the silanization. Such polymer segment
confinement would lead to a hindered polymer dynamics within nano-scale
ranges, playing a key role in the frequency domains associated with either
wet skid or rolling resistance. Further works are needed to fully support such
proposals, but recently published data document indeed the very special
dynamic–mechanical properties of silica-filled S-SBR compounds, with very
small effects assigned to silanization (with TESPT) up to 70 phr silica.

6.1.1.5  Silica and Polydimethylsiloxane


Polyorganosiloxanes, the so-called “silicones,” whose general formula is
R′–(SiOR2)n–R′, are a family of polymer materials with unique and interest-
ing properties.3 Silicone polymers have the alternating –Si–O– type structure
as part of their backbone chain and although silicon is in the same group
as carbon in the periodic table, it has quite a different chemistry. Various
Polymers and White Fillers 247

organic groups, e.g., methyl, benzyl, etc., may be bonded to the silicon to
yield polymer materials that are by nature water repellent, heat stable, and
very resistant to chemical attack. Poly(dimethylsiloxane) or PDMS is by far
the most common silicone polymer, whose flexibility is due to the inorganic
siloxane backbone, with a very low surface energy imparted by the methyl
groups. This results in a glass-transition temperature of less than –120ºC,
and consequently quite a large usage temperatures window, from below
–40ºC to above 150ºC.
Various PDMS materials are obtained through synthesis and hydrolysis of
chlorosilanes, then polycondensation, according to the following schema:

Si + 2 ClCH3 → SiCl2(CH3)2

n SiCl2(CH3)2 + 2 n H2O → Si(OH)2(CH3)2 + 2 n HCl

n Si(OH)2(CH3)2 → [SiO(CH3)2]n + n H2O

The value of n fixes the molecular weight, and hence the viscoelastic nature
of the material, from low viscosity oils up to high MW polymers. Vulcanizable
elastomers are obtained by introducing reactive sites, for instance vinyl
groups. Mechanical properties of silicone polymers are improved through
the addition of fillers, with silica the most obvious choice. Globally the same
reinforcing effects as in other elastomers are observed with silicone/silica
compounds, with the level and the structure of the filler playing qualitatively
the same roles. However there are a few singular aspects in polysiloxane/
silica systems worth discussing in details because certain well established
scientific knowledge can be somewhat extrapolated to other systems.
Interactions between organosiloxanes and silica particles, either fumed or
precipitated is a long studied subject, either for purposely promoting graft-
ing chemical reactions on the particles23 or as an approach for understand-
ing the interactions between siloxane and silanol groups.24 One would a
priori consider that, owing to their close chemistry, polysiloxanes and silica
are naturally compatible and that it is relatively easy to disperse the latter
in the former. This is globally true but it has been observed for long that,
once a silica and a silicone polymer have been mechanically mixed, the very
adsorption process of polymer chains on the surface of silica particles is rela-
tively slow, even if it tends to accelerate at higher temperature. For instance,
at 70°C, three months are necessary for a PDMS sample to fully saturate the
silica surface. The adsorption properties and kinetics in silica/PDMS sys-
tems, the structure of the adsorbed layer and other singular aspects were
investigated by a number of authors and their findings allow to somewhat
understand certain engineering practices that were pragmatically developed
by manufacturers of silicone products, e.g., mastics, sealants, and other self-
­vulcanizing liquid silicones. Early observations revealed that in useful silica/
248 Filled Polymers

silicone dispersions, the polymer on the silica substrate is approximately


ten times more concentrated than what would correspond to a monolayer
of water.25 Then, by combining chemical microanalysis, Nuclear Magnetic
Resonance (NMR) relaxation and swelling measurements, it was shown that
directly after mechanical mixing, PDMS chains are strongly adsorbed on
only a quarter of the silica surface, in the form of adsorbed islets.26 It was
first established that, during the slow saturation of silica surface, which lasts
around three months at 70°C, the average number of contact points of a chain
with the silica surface is proportional to N b (Nb is the number of skeletal
bonds in one chain).27 Eventually using NMR relaxation and diffusion stud-
ies, the structure of silica/PDMS systems was rationalized by a three-state
model: close to silica particle, there is a (strongly) bound polymer layer, then
polymer chains are entangled or restricted by the adsorbed layer and even-
tually there is the free bulk polymer.28 Recent works with modern sophisti-
cated analytical techniques essentially confirm previous observations.29,30 It
is worth noting here the remarkable similarity between such a description of
the structure of silica/silicone systems and the established morphology for
carbon black filled rubber compounds (refer to Figure 5.16). Thanks to their
similar chemistry, silica and silicone polymers develop thus instant interac-
tions but the full capabilities of such systems are obtained only when the
structure of the adsorbed polymer is completely developed.
The adsorption kinetics is consequently an issue of industrial importance
with silica/silicone systems and was consequently thoroughly investigated.
Cohen-Addad et al.31 studied the time dependence of the adsorption of
siloxane chains on silica aggregates and proposed the following empirical
model:

[Qsat − Q(t)] = [Qsat − Q0 ] × exp  − t 


tad 
(6.1)

where Q(t) is the amount of bound polymer at time t (in g/g of silica), Q 0
and Qsat respectively, the initial (i.e., directly after mechanical mixing) and
final (i.e., at saturation) amounts of adsorbed polymer (in g/g of silica), t the
time and tad a characteristic time. Experiments at room temperature with
either hydroxyl or methyl terminated polymers and 29%wt silica gave very
high values for tad, i.e., 7.1 × 107 s (around 2.25 years) and 6.2 × 108 s (nearly 20
years), respectively. Whatever their chain end, polydimethylsiloxanes adsorb
on silica particles according to simple equivalent laws of adsorption, propor-
tionally to Mn .32 For a given silicone polymer, there is a specific silica con-
centration for a tri-dimensional silica-polymer morphology to be obtained,
with the associated viscoelastic behavior (i.e., gel or no-gel).33
Equation 6.1 is essentially an empirical model, likely selected for its con-
venience in fitting experimental data, using a linear algorithm for instance
the least square method. It is worth noting that considering the variation of
Polymers and White Fillers 249

the adsorbed amount of polymer with respect to the square root of time is
an obvious reference to a Fickean process. Providing nonlinear fitting algo-
rithms are available, an equivalent but more explicit model is:

Q ( t ) = Q0 + Qinf 1 − exp ( − λ t ) (6.2)

where Q 0 is the initial adsorbed polymer (i.e., directly after mechanical mix-
ing) and Qinf the additionally adsorbed polymer after an infinite time (i.e.,
at saturation), both in g/g of silica), t the time and λ is a parameter related
with the characteristic time tad for the adsorption process (i.e. λ = t-0.5
ad). The
overall bound polymer for an infinite time is then given by (Q 0 + Qinf), which
corresponds obviously to Qsat in Equation 6.1. As shown in Figure 6.8, experi-
mental data on silica/PDMS systems34 were fitted with Equation 6.2, using a
nonlinear fitting algorithm (see details in Appendix 6.1).
As can be seen the model meets well experimental data and the initial
and the final quantities of polymer are directly obtained as fit parameters.
The initial quantity of adsorbed polymer is somewhat depending on the
mechanical mixing conditions (unfortunately not precisely documented
in the source of data) but it is quite clear that, at equal silica content, the
higher the molar mass of the polymer, the larger the bound polymer content
at the end of the mixing step. Increasing the silica loading gives expectedly a
higher initial bound polymer but also a lower quantity of adsorbed polymer
at saturation.
The value of the characteristic time tad (or the parameter λ in Equation 6.2)
gives an insight on the time scale of the adsorption process. Even if increas-
ing the temperature somewhat speeds up the process, quite long maturation
periods are necessary for the silica surface to be fully saturated. As clearly
seen in Figure 6.8, higher Mn polymers tend to mature faster and silicone
product manufacturers obviously take advantage of this effect in tailoring
their products. But slow maturation processes mean also that some “age-
ing” either on storage, or during the life of the material can be expected, as a
mere result of polymer chain dynamics in the vicinity of silica particles. For
instance, DeGroot and Macosko investigated the aging behavior of silica/
PDMS systems, by measuring the rheological properties, the bound rubber,
and the state of dispersion as a function of time.35 They observed softening
rather than hardening as typically reported for silica filled systems. Polymer
adsorption onto the surface plays obviously an important role in determin-
ing the overall stability of these systems and the addition of a surface treat-
ing agent (e.g., hexamethyldisiloxane and hexamethyldisilazane), either
physically adsorbed or covalently bound to the silica surface, can inhibit
the adsorption of polymer. A better stability of the silica dispersion is then
observed and therefore variation in rheological properties are reduced.
With respect to the silica surface chemistry (see Figure 6.1) and the chemical
nature of polysiloxane, the interaction sites are clearly identified since they
250

Filler weight fraction = 0.20 PDMS Mn = 43000 g/mol


2.5 2
Φ = 0.049
2
Mn = 300,000 g/mol 1.5

1.5
Φ = 0.103
Mn = 73,000 g/mol 1
1
Φ = 0.204
0.5
0.5 Mn = 43,000 g/mol

Adsorbed polymer Q (g/g of filler)


Adsorbed polymer Q (g/g of filler)
0 0
0 1000 2000 3000 4000 0 1000 2000 3000 4000
Time (h) Time (h)
Mn, g/mole : 43000 73000 300000 Silica fraction : 0.049 0.103 0.204
Q0 (g/g) 0.257 0.320 0.733 Q0 (g/g) 0.103 0.266 0.255
Qinf (g/g) 0.588 1.321 1.483 Qinf (g/g) 1.702 1.042 0.614
λ 0.044 0.053 0.061 λ 0.041 0.022 0.039
tad (h) 514.8 350.7 273.1 tad (h) 588.7 2020 648.2
r2 0.991 0.987 0.992 r2 0.980 0.993 0.985

Figure 6.8
Adsorption kinetics of polydimethylsiloxane on silica particles. (Experimental data from L. Dujourdy, PhD Thesis, University of Grenoble, France, 1996;
nonlinear fitting of Equation 6.2.)
Filled Polymers
Polymers and White Fillers 251

essentially concern hydrogen bondings, either at the ends of the polymer


chain (if they are hydroxylated) or through any of the –O– of the chain. It fol-
lows that the area of an interaction site is perfectly known, by calculating the
“area” of one silanol group, around 0.55 nm2 (55.10–20 m2). In comparison with
carbon black/elastomer systems, having a clear identification of the filler-
matrix interaction sites is a tremendous advantage in developing a theoreti-
cal approach of the adsorption process. Indeed, starting from the percolation
theory with only two assumptions, Cohen-Addad developed a model for the
bound PDMS polymer on silica at saturation.36 The assumptions are (1) that
PDMS chains obey Gaussian statistics, (2) that there is hydrogen bonding at
each PDMS-silica contact point. The model is written as follows:

M0 c Sp
Qsat = ⋅ ⋅ Mn (6.3)
A0 ε 0 N Av

where Qsat is the bound polymer at saturation (g/g of silica), M0 the mass of
the monomer unit [75 g/mole for -Si(CH3)2-O-], A0 the area of one interaction
site (i.e., 0.55 nm2), c the filler concentration (g of filler/g of polymer), Sp the
specific surface of silica, NAv the Avogadro number, Mn the number average
molar mass of the polymer and ε0 a factor for the stiffness of the chain ( ≈ 1
in first approximation). Figure 6.9 shows how experimental data34 on various

1
[1 – exp(– 0.065 √Mn × ΦSil)]

0.8

0.6
Qsat × ΦSil

0.4
Ssp (m2/g)

150
0.2 50
300

0
0 100 200 300
√Mn × ΦSil

Figure 6.9
Adsorption of polydimethylsiloxane on silica particles; Cohen–Addad percolation model
for maximum adsorbed polymer at saturation (Equation 6.3) vs. experimental data from L.
Dujourdy PhD Thesis, University of Grenoble, France, 1996.
252 Filled Polymers

silica/PDMS systems, largely varying in polymer molar masses and/or silica


contents, fall on a single curve when calculating the maximum adsorbed
polymer at saturation and suitably reducing the scales.
Having the interaction site well identified in a filled polymer system, in
terms of chemical activity and surface, and a clear picture of the nature of
the polymer–filler interaction allow quite convincing theoretical models to
be developed. Such a favorable situation is however restricted to a few cases,
namely silica/polysiloxane systems. With other systems, either the nature of
the polymer–filler interaction is badly known or the size of the interaction
site cannot be clearly quantified, or both. In such case however, the success-
ful silica/PDMS case provides some interesting guidelines when assuming
that, whatever are the respective chemical natures of the filler and the poly-
mer, at least the physics is the same. As we have seen the author has success-
fully adapted this model to the case of carbon black/rubber systems, with
however the additional difficulty that the surface area of the interaction site
A0 cannot be known a priori (see Chapter 5, Section 5.1.5).
Owing to the interactions described above, gels are easily obtained when
adding silica particles to liquid PDMS, and without further cross-linking,
such gels find niche applications such as protective materials for fiber-optic
cables and as encapsulants for semiconductor devices. As may be expected,
the rheological and mechanical properties of silica-PDMS gels are quite
complex, namely with respect to the viscoelastic behavior, but are well doc-
umented in the research literature, likely because it is relatively easy to pre-
pared silica filled polysiloxane systems with standard laboratory equipment.
Hysteretic effects upon increasing shear rate in both the viscosity and the
first normal stress difference, as well as significant overshoot in the stress
growth function, were reported by Caruthers and colleagues,37,38 and inter-
preted in terms of interparticles interactions via entanglements of the poly-
mer adsorbed on the (fumed) silica surface. After an applied shear is step
changed, the shear stress relaxes or grows in a complex manner that depends
on the shear history. Consequently, PDMS molecular weight and silica vol-
ume fraction play an important role in such effects, as well as the surface
chemistry when modified in the appropriate manner, but the preparation
procedures (mixing technique and time) and the sample age further affects
the rheological behavior of such systems.39–42 Precipitated silica develop also
strong interactions with PDMS, and bring similar effects, but less than that of
fumed silica.43 At high volume concentrations of precipitated silica (0.128 and
0.160), a yield behavior is evident from the storage modulus measurements.
It is worth underlining that the key for understanding all those effects is the
physical adsorption of PDMS on the silica surface, without chemical bond-
ing, as clearly demonstrated by several authors.44,45
Suitably cross-linked, an unfilled, high molecular weight polydimethyl-
siloxane exhibits very modest mechanical properties, for instance a tensile
strength (TS) in the 0.35 MPa range, largely insufficient for most applications.
But the addition of a reinforcing filler, such as a high structure silica, increases
Polymers and White Fillers 253

the same property by a factor of around 40, yielding products with TS in the
13–14 MPa range, and around 600–700% elongation at break. Such reinforc-
ing effects are by large more important than what is currently achieved with
reinforcing fillers (e.g., carbon black and silica) in common hydrocarbon elas-
tomers. This unusually high degree of reinforcement observed with silica/
polysiloxane systems has long been attributed to the particular polymer–
filler interactions previously described, which persist after exposure to high
temperature curing.46 No chemical bonding has been demonstrated between
silica and (uncrosslinkable) polysiloxane but, in a cured PDMS-silica system,
one cannot exclude a combination of chemical and physical bonds. The for-
mer are likely covalent bonding occurring upon vulcanization; the latter are
hydrogen bonding and van der Waals forces, indeed favored by a high struc-
ture of filler particles. There are of course a variety of finely divided miner-
als that can be used as fillers for (curable) polysiloxanes, for instance finely
grinded quartz, or zinc, titanium and iron oxides, or calcium carbonate, but
amorphous silica in the 150–400 m2/g surface area range provides the best
reinforcement. In order to have easy-to-process systems, it is however neces-
sary to prevent certain detrimental polymer–filler interaction prior to vul-
canization by using suitable plasticizers, for instance low molecular weight
polysiloxane oligomers.
Figure 6.10  shows typical dynamic properties of vulcanized PDMS-
silica systems, as investigated through strain sweep experiments at constant
frequency and temperature.47 As can be seen, dynamic strain softening is
observed in a qualitatively similar manner to other filled polymers. It fol-
lows that models, which successfully fit conventional filled rubbers (e.g.,
carbon black filled compounds), are expected to well suit such data. This is
indeed the case, as shown by the curves in Figure 6.10, drawn by fitting the
Kraus–Ulmer equations, i.e.,

G0′ − G′inf
G′ ( γ 0 ) = G′inf + 2m
(6.4a)
γ 
1+  0
 γc 

and

m
γ 
′′ )  0 
2 (Gm′′ − Ginf
 γc   γ 
G′′ ( γ 0 ) = Ginf
′′ + + Gk′′ exp  − 0  (6.4b)
γ 
2m
 γk 
1+  0
 γc 

The fitting parameters are given in Table 6.3. In agreement with the physi-
cal reasoning that supports the model, the critical strain γc decreases with
254

PDMS + Silica; 1Hz; 25°C Filler fraction


10 0.04 10
0.08
8 0.12 8
0.15
6 0.18 6
0.21
4 4

2 2

Elastic modulus G´, MPa


Viscous modulus G´´, MPa
0 0
10–4 10–3 10–2 10–1 100 10+1 10–4 10–3 10–2 10–1 100 10+1
Strain amplitude Strain amplitude

Figure 6.10
Dynamic strain softening as observed on PDMS-silica systems; 0.078% vinyl-pendant PDMS (Mn=140,000 g/mol; Mw=390.000 g/mol); 300 m2/g silica;
peroxide crosslinked. (Experimental data from L. Dujourdy, PhD Thesis, University of Grenoble, France, 1996.)
Filled Polymers
Polymers and White Fillers 255

Table 6.3
Modeling the Dynamic Strain Softening Effect on PDMS-Silica Systems
with the Kraus–Ulmer Equations
Φsilica 0.04 0.08 0.12 0.15 0.18 0.21

Modeling G′ vs. Strain Amplitude (Equation 6.4a)


G′inf 0.038 0.403 0.327 0.236 0.616 1.010
G′0 0.457 0.698 1.459 3.008 4.423 8.252
γc 0.3398 0.0740 0.0392 0.0346 0.0235 0.0195
m 0.144 0.671 0.284 0.229 0.308 0.385
r2 0.9872 0.9969 0.9979 0.9985 0.9986 0.9990

Modeling G′′ vs. Strain Amplitude (Equation 6.4b)


G″inf (–0.108) 0.200 (–0.023) 0.451 1.023 2.780
G″m 0.106 0.301 1.046 2.523 4.224 8.010
γc 0.0382 0.0063 0.0415 0.0302 0.0288 0.0300
m 0.201 0.100 0.295 0.417 0.476 0.611
G”k 0.071 0.295 0.624 0.762 1.000 1.154
γk 0.0022 0.0001 0.0009 0.0010 0.0011 0.0021
r2 0.8642 0.7777 0.9504 0.9961 0.9966 0.9990
Note: Fit parameters; note that negative values for G″inf have no physical meaning,
likely reflect experimental inaccuracy and could be replaced by very low posi-
tive numbers.

increasing filler fraction and is not much different for G′ and G′′. However
the exponent m is typically depending on the silica content and is nearly
twice as large for G′′. The additional Ulmer term in Equation 6.4b allows to
meet the G′′ vs. γ behavior at low strain, with the main result that the critical
viscous modulus G′′k significantly increases with the filler fraction.
The above data allows however to demonstrate how really strong are the
PDMS-silica interactions. Indeed, using the fit parameters in Table 6.3 and
the Kraus-Ulmer equations, one easily calculates low strain (let’s say 0.001)
values of G′ and G′′, in order to draw Figure 6.11.
Since the dynamic properties of the pure polymer were not given in the
source data, G′(Φ = 0) and G′′(Φ = 0) were obtained by second degree extrap-
olation. The left graph shows that 20% silica increase the elastic and viscous
moduli by a factor of respectively, 42 and 95. But because the elastic modulus
of the pure polymer is considerably larger than the viscous modulus, filled
materials still exhibit a strong viscoelastic character. In the right graph, the
normalized complex modulus is plotted vs. the filler fraction. The complex
modulus is calculated as G * = G′ 2 + G′′ 2 and normalized with respect to
G * of the pure, unfilled material. In other terms, one plots the functional
for the silica effect, in order to compare it with the well known Guth and
Gold term for filled systems, when only hydrodynamic effects occur, i.e.,
(1 + 2.5 Φ + 14.1 Φ2). Of course, when significant interactions exist between a
256

PDMS + Silica; 1 Hz; γ = 0.001; 25°C


8 50
7 : G´
: G´´ 40
6

5
30
4
20

G*(Φ)/G*(0)
3

at 0.001 strain, Pa
2

Dynamic moduli G´ and G´´


10

Normalized complex modulus


1 Guth & gold
0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Filler fraction Φ Filler fraction Φ

Figure 6.11
PDMS-silica systems; variation of the low strain (γ = 0.001) dynamic properties with filler fraction; the dash curve in the right graph is the Guth and
Gold term for mere hydrodynamic effects.
Filled Polymers
Polymers and White Fillers 257

polymer matrix and dispersed mineral particles, one does not expect the
Guth and Gold model to meet experimental data, but the right graph in
Figure 6.11 is an impressive (and simple) demonstration of the exceptional
reinforcing effect of high structure silica in polysiloxanes. As we will see
later, other (white) fillers are far to give effects of such a magnitude.

6.1.2 Elastomers and Clays (Kaolins)


Kaolins have the finest particle-size range of all naturally occurring white
fillers. Compared to (silanated) silica, kaolins or clays have relatively mild
reinforcing properties but satisfactory however in a number of applications,
for which specific effects are obtained owing to their platy structure. Clays in
rubber allow very high hardness parts to be fabricated, essentially because
of their plate-like particles, in contrast with spherical particles of equivalent
dimensions. So-called “hard clay” grades have reinforcing capabilities, i.e.,
higher modulus, tensile strength and resistance to abrasion. “Soft clays”
give lower physical properties and therefore are rather extenders.48 There
are essentially four qualities of kaolin, depending on the recovery process
and/or the after mining treatment: air-floated, water-washed, delaminated,
or calcined. Airfloat hard clay is a general purpose white extender in rubber
applications. Other grades are used for specific purposes, e.g., (clear) color
enhancement, better mechanical properties and  abrasion resistance for
waterwashed and delaminated clays. Calcined and surface-treated clays are
used for improved electrical properties in low and medium voltage power
cables, with the insulation properties maintained under wet operating con-
ditions owing to the low moisture adsorption of the calcined and coated
particles. Rubber extrusion and calendering generally benefit from the pro-
cessing aid capabilities of calcined clays, so that hoses, profiles, and sheets
can be manufactured with a very smooth finish and an excellent dimen-
sional stability. With certain specialty rubbers, e.g., polychloroprene, butyl
rubber and chlorinated polyethylene, calcined kaolins give very low levels
of mill sticking. Light colored and chemically inert, calcined clays allow to
prepare compounds for applications where only high quality materials are
permitted, for instance pharmaceutical closure applications.
There are many grades of kaolins available worldwide, certain only locally.
As illustrated in Table 6.4, it is not easy to compare the various products on
the market, essentially because there is no standard description of such
materials. Most suppliers give (in their published technical data sheets) a
limited number of common information that would allow some compari-
son. Specific gravity is the same whatever the origin, essentially because
the mineral composition is fairly constant, approximately 45–55%wt SiO2,
35–41 Al2O3 and smaller quantities of various oxides, certain obviously
affecting the degree of whiteness (e.g., Fe203). In what the mechanical and
rheological properties are concerned, the exact mineral composition of
kaolin is not a relevant information; it is clearly the particle size and the
Table 6.4
258

Selected Commercial Kaolin Grades from Various Suppliers


Specific Gravity Specific Surface Area, Oil Absorption Particle size (Distribution)
Producer Brand Name (g/cm3) BET (m2/g) (g/100 g) (µm)
IMERYS PoleStarTM 200R 2.6 8.5 n.a. 44% < 2
(Worldwide) Calcined clay
PoleStarTM 450 2.62 n.a. 60 1.5 (median)
Calcined clay
SpeswhiteTM 2.6 14 42 0.5 > 10
Ultrafine powder 80% < 2
SupremeTM 2.6 n.a. n.a. 0.2 > 10
Powdered 77% < 1
GlomaxTM LL 2.62 9 53 1.5 (median)
High temperature treated
SERINA Filler grade 2.6 18.73 45–50 98% < 20
(South Africa) Water washed 90% < 10
48% < 2
Ultrafine 2.6 21.25 57–63 99% < 20
Water washed, 98% < 10
Screened 70% < 2
VANDERBILT Dixie Clay® 2.62 n.a 42 99.3% < 78
(USA) (Hard clay) 0.2 µm (median)
Par® Clay 2.62 n.a. 40 99.3% < 78
(Hard clay) 0.2 µm (median)
Filled Polymers
LangfordTM 2.62 n.a. 36 99.0% < 78 h
(Soft clay) 1.3 mm (median)
Peerless® 1 2.62 n.a. 30 99.6% < 78
Polymers and White Fillers

1.2 mm (median) 30
SOKA Sialite 2.6 9 41 99% < 20
(France) Powdered 85% < 10
63% < 5
38% < 2
Blankalite 78 2.6 15 36 99% < 20
Powdered 97% < 10
76% < 5
51% < 2
Metasial V800 2.6 6–7 48 100% < 20
Calcined clay 96% < 10
70% < 5
35% < 2
259
260 Filled Polymers

particle size distribution that are the key characteristics,49 unfortunately


not documented in a comparable manner by the various suppliers.

6.1.3 Elastomers and Talc


Ground talc can be considered as a functional additive for elastomers, because
its specific structure and aspect ratio may provide interesting enhancements
of certain properties, namely mechanical properties, i.e., modulus, tensile
strength and elongation at break, hot tear resistance, useful for easy demould-
ing parts with complex geometries, and barrier properties. In tire technology,
the use of talc in innerliner and white side-wall compounding (the latter appli-
cation no-longer of fashion today) has been reported.48 Talc (and other platy
fillers) is known for long to restrict the swelling of rubber parts exposed to
automotive fuels, oils and greases, an effect essentially due the correct orienta-
tion of the platy particles, in such a manner that a strong swelling anisometry
is obtained, because of obvious barrier effects. Indeed talc particles have large
planar surfaces that are hydrophobic and thus preferentially wetted by organic
substances such as elastomers. Talc incorporation in elastomers is thus easy
and, owing to their aspect ratio, talc platelets easily self-orient along flow lines.
This give talc the role of a processing aid through internal lubricating effects
between elastomer chains. Easier extrusion is therefore obtained, as well as
reduced extrudate swell. Talc’s lamellar structure give barrier effects, substan-
tially reducing air and fluid permeation in stopper, sealing, membrane and
hoses; improved thermal and weathering resistance is also obtained through
similar barrier effects. Talc platelets are electrically neutral, so that interesting
insulation effects are obtained in wires and cables.
Oriented platy fillers provide elastomers quite unique failure mechanisms
that, properly exploited, give materials with an enhanced resistance to destruc-
tive crack growth. For instance Eldred compared chlorosulfonated polyeth-
ylene compounds with equal volume fraction of either N550 carbon black or
6 µm platy talc particles.50 Through the appropriate procedure on a two-roll
laboratory mill, samples where prepared with a purposely anisotropy. Using
results from trouser tear, peel and De Mattia fatigue tests, Eldred clearly dem-
onstrated that talc filled compounds failed in an interlaminar shear mode,
parallel to the applied stress, not resulting however from planes of weakness
in the lateral direction, but from a more than five-time increase in the cohesive
fracture energy with respect to the same measurements on the carbon black
filled control compound. Fractured surfaces observed by scanning electron
microscopy reveal that this enhanced crack propagation resistance was due
to a more efficient fracture energy absorption, as resulting from a diversion
of the fracture paths through the elastomer, owing to the presence of the talc
particles. Such results are important because they clearly demonstrate the
strong links to (may) exist between the particular rheology imparted to com-
posites by platy fillers and the resulting benefits in mechanical properties,
providing an adequate control of flow induced anisotropy effects is achieved.
Polymers and White Fillers 261

Extruded or calandered products are obvious practical applications for such


effects. There are however some negative aspects which limit the reinforcing
capabilities of talc in tire applications. Mouri and Akutagawa51 reported that
when substituting 20 phr of N234 carbon black in a typical OESBR/BR tread
compound by 20 phr talc (with also a 20–5 phr reduction in process oil con-
tent), better wet traction and lower rolling resistance are obtained but with
a significant reduction in wear resistance. It must be noted however that the
grade of talc used by these authors was considerably coarser than N234 (par-
ticle size in the 200–8000 nm range for the former, vs. 17–26 nm for the black)
and when more balanced formulation changes are made with an ultra-fine
talc grade, substantially different results are obtained, as recently reported by
Noel and Meli.52 These authors showed by adding only 5 phr of an ultra-fine
talc (Mistron® Vapor R; Luzenac, Rio Tinto Group) to either N220 black filled
NR/BR or Silica filled SSBR/BR compounds, significant improvement in filler
dispersion were obtained, at equal or even 20% reduced mixing time. Not
significant modification were seen in ultimate tensile, dynamic and adhesion
properties. Only a very slight decrease in wear resistance was noted with
respect to the carbon black filled compound, with however a beneficial effect
on De Mattia fatigue resistance. Some possible interferences with the silica
silanization process were also noted.
It is well known that the basal surfaces talc platelets are hydrophobic, while
the edge surfaces are hydrophilic.53 The hydrophobic character of the basal
surfaces arises from the fact that the atoms exposed on the surface are linked
together by siloxane (Si-O-Si) bonds and therefore cannot develop hydrogen
bonding with water. The edge surfaces, however, are made of hydroxyl ions,
Mg, Si and substituted cations; all of which may undergo hydrolysis. It fol-
lows that the edges are hydrophilic, and can form strong hydrogen bonds
with water and/or polar substances. Chemical modification of talc with
appropriate chemicals is therefore possible, in order to optimize its com-
patibility with the polymer matrix, through possibly (and ideally) chemical
adhesion. As there is a limit (either technical or economical) in reducing the
size of any naturally occurring mineral, adding any type of organic chemi-
cal is a common method to produce so-called surface-treated minerals. But
simple surface coating is substantially different from surface modification
based on reactive functionality available on the mineral particles.
A truly surface-modified talc has a coupling agent durably attached by
covalent bonds and this can only be achieved through the hydroxyl groups
available at the edges of the platelets. Depending if the modifying agent is bi-
functional or not, chemical bonding with the polymer could eventually occur
through the appropriate chemical reaction or at least chain entanglement
may be expected. The most commonly used talc modifiers are the organosi-
lanes, whose general chemical structure is RSiX3, where X is a hydrolysable
group, such as methoxy, ethoxy, acetoxy, or chloride, and R is a nonhydrolys-
able organofunctional group. The most often used silanes are alkoxy deriva-
tives, RSi(OR’)3, with amino, epoxy, methacrylate or vinyl functionality.
262 Filled Polymers

As we have seen, silanization has been most studied in the case of silica
which are considerably richer in surface hydroxyl groups, so-called silanols,
than talc. Because the effectiveness of silane modification depends on the
availability of silanol groups on the mineral surface, some of the silicate fill-
ers are more amenable to silane modification than others. Hydroxyl groups
are available on the edge of talc platelets, so that silane treatment is feasible
and, indeed, silane treated talc grades are commercially available. Silane
treatment levels are typically 0.5–1.0% on mineral weight, to be compared
with the 8% recommended for the silanization of synthetic silica.
Whilst not comparable with what can be achieved with (silanated) silica,
very fine microcrystalline talc grades offer quite attractive reinforcing prop-
erties for white rubber compounds, particularly when a pretreatment with a
suitable organo-silane is used. Not much published data is available regard-
ing the (industrial) silane modification of talc (as well as mica and kaolin) but
commercially available silane-modified grades offer proved technical ben-
efits when used as fillers or additives to polymer systems. With respect to the
general chemical structure of the most widely used organo-silanes, it is per-
mitted that industrial modification of talc involves the following four steps.
First, hydrolysis of the three labile groups must occur, then condensation to
oligomers follows. The oligomers then hydrogen bond with OH groups at
the edges of talc platelets. It is eventually during drying or a further ther-
mal treatment that a covalent bond linkage is formed, with a concomitant
loss of water. Such reactions are likely to occur simultaneously after the ini-
tial hydrolysis step. The relatively low quantities of (supposedly) chemically
bound silane appear sufficient to drastically modify certain key properties of
the filler, namely its dispersibility in diene rubbers and hence the associated
rheological and mechanical benefits.
It is worth mentioning at last an important use of talc in the rubber
­industry, whilst not as a filler. Relatively coarse grades of talc (average parti-
cle size ≈ 40–50 µm) are used as partitioning agent, either in powder form or
as suspension in water, to coat freshly compounded, uncured rubber strips
and sheets, in order to avoid stickiness during storage, before further process-
ing operations. Surface coated talc is easily incorporated in the compound
during subsequent shaping steps, without generally undesirable effects.

6.2  Thermoplastics and White Fillers


6.2.1 Generalities
Most of the minerals used as fillers in thermoplastics are “white” with how-
ever a great variety in chemical nature and/or physical forms. Generally fill-
ers are added to thermoplastics for economical reasons with possibly some
Polymers and White Fillers 263

benefits in mechanical or physical properties. Notable exceptions (i.e., where


cost is not the prime issue) are when improvements in certain specific prop-
erties of thermoplastics are sought, for instance fire resistance or flame retar-
dancy through the addition of finely grinded aluminium tri-hydrate (ATH),
Al2O3.3H2O, magnesium hydroxide Mg(OH)2 or antimony tri-oxide Sb2O3. If
cost saving is the main reason, then two problems appear: 1. losses (or at
least undesired changes) in mechanical properties must be minimized; 2.
the rheology and hence the processing behavior of filled compounds must
be mastered and/or controlled.
Besides cost, white fillers globally affect the following properties of
thermoplastics:

• Stiffness and hardness generally increase with higher filler content


• Impact resistance is either decreased or improved depending on the
filler nature and its characteritics, namely particle size
• Thermal expansion and shrinkage are reduced
• Softening temperature (Vicat, HDT) is increased
• Flow properties are modified and either easier or worst processing
behavior is obtained
• Surface properties (gloss, scratch resistance, etc.) are modified
• Fire resistance is improved

But the effect of a given (type of) filler is frequently specific to a polymer
or to a class of polymers and, quite often, there are antagonistic effects. For
instance, benefits in mechanical properties are counterbalanced by increas-
ing processing difficulties, or the fire resistance imparted by the additive
(e.g., ATH in polyester or epoxy resins) goes along with an excessive increase
in viscosity. A priori, the chemical compatibility between filler surface and
polymer segments is critically important in both the wetting and dispersion
of particles in the matrix, the processing behavior of molten composites and
eventually the (mechanical and physical) performances of final objects. But
except in a few cases, the actual surface chemistry of many white fillers is ill-
known, and at best referred in terms of affinity for water, i.e., hydrophilic or
hydrophobic material. Hydrophilic fillers would give maximum interaction
with polar polymers, hydrophobic fillers would be preferred with nonpo-
lar polymers. It follows that many commercial fillers, especially of mineral
origin, are surface coated or chemically treated with hydrophobic wetting
agents, in order to modify their surface chemistry, or at least to alter their
wetting character. Dispersion in nonpolar polymers is expectedly easier,
likely because wetting agents help in deagglomerating clustered filler par-
ticles during the mixing process. This aspect is particularly important for
continuous mixing operations for which the viscosity of the molten com-
posites must be kept low, at levels compatible with the performance of the
264 Filled Polymers

equipment. Typical wetting agents are fatty acids and derivatives, polymeric
esters, and organosilanes, the latter widely used with respect to their reactive
potential with suitable functional groups on the filler surface. Except silica,
whose surface is known to be silanols rich, the real (i.e., chemical) effect of
silane treatment of many white mineral fillers is either unproved or unclear
in the author’s opinion.
Table 6.5 illustrates the usages of a few selected white fillers in thermoplas-
tic polymers.
Table 6.5
Typical Industrial uses of a Few Selected Mineral Fillers
General Purpose Loading
Fillers Polymer Range (%) Improvements/Effects/Application
Calcium PP, PS, PVC, ABS, 5–80 Rheological properties
carbonate POs, TPE, PU, Mechanical properties
(grinded or Epoxy, Phenolics, Surface aspect
precipitated) Fluoropolymers Numerous applications
Clays, silicates, POs, PU, PVC, PA, 10–40 Rheological properties
kaolin TPE Surface aspect
Wire and cables
Automotive parts
Mica PP, ABS, POs, PC 5–40 Mechanical properties
TPE, PA, Barrier effects
Fluoropolymers Dielectric properties
Thermal properties
Silica ABS, POs, PS, 10–30 Rheological properties (thixotropy)
PVC, PU, TPE, Extender or thickener
Epoxy Reinforcement
Talc PP, POs, PVC, PS, 20–50 Extender
Phenolics Reinforcement (stiffness, tensile)
Creep resistance
Barrier effects
Loading
Specialty Fillers Polymer Range (%) Improvements/Effects/Application
Aluminum ABS, LDPE, PVC, 9–50 Extender
tri-hydrate TPE Flame retardant
Smoke suppressant
Wire and cable
Barium sulfate PU 10–30 Specific gravity increase
(barites) Surface properties (friction)
Chemical resistance
Wollastonite PC, PS, TPES, PA 20–50 Mechanical properties (tensile)
Dimensional stability
Barrier effects
Electrical properties
Thermal properties
Note: PP: polypropylene; PS: polystyrene; PVC: poly(vinyl chloride), POs: polyolefins;
TPE: ­thermoplastic elastomers; PU: polyurethane; PA: polyamides; ABS: acrylonitrile-
butadiene-styrene terpolymer; LDPE: low density polyethylene.
Polymers and White Fillers 265

With respect to mechanical properties, the effect of any mineral filler is


firstly related to particle size and particle size distribution. For instance,
the impact resistance of a PP–talc composite strongly depends on those two
filler characteristics, and it is easily understood that particles with exces-
sive dimensions are likely to be fracture initiation sites. For mineral fillers
which are extracted from lodes, grinding and sorting are thus key prepara-
tion steps, with obviously processing costs that dramatically increase as the
targeted (average) particle size decreases. The balance between the benefits
in mechanical or physical properties and the extra-cost of filler addition is
therefore the bottleneck of all applications for filled thermoplastics.
Besides the (beneficial) effects on flow and mechanical properties, not all
fine white minerals can be considered as filler for thermoplastics because,
depending on their hardness, some can bring excessive wear of processing
equipment. It follows that the Mohs hardness is a relevant property in select-
ing (white) mineral fillers, notwithstanding their other effects. The Mohs
scale is one of several definitions of hardness in materials science and con-
veniently characterizes the scratch resistance of a given mineral through its
ability to scratch a softer one. It was established in 1812 by the German min-
eralogist Friedrich Mohs (1773–1839) using ten minerals, all readily avail-
able. Diamond, the hardest known naturally occurring substance, is at the
top of the scale, and talc at the bottom. The hardness of a given material is
measured against the scale by finding the hardest material it can scratch,
and/or the softest material that can scratch the given material. For instance a
material that is scratched by calcite but not by gypsum has a Mohs hardness
between 2 and 3. The Mohs scale is a so-called ordinal or successive scale,
thus somewhat arbitrary, and does not measure the actual hardness. The
Mohs scale is roughly linear with respect to the logarithm of the absolute
hardness, as measured with a “sclerometer,” a special instrument invented
in 1896 by T. Turner and used by mineralogists. It consists essentially in mea-
suring with a microscope the width of the scratch made by a diamond when
drawn under fixed load across the face of the specimen under test. The wider
the scratch the softer the material and the results are expressed with respect
to an increasing scale, starting from 1 (talc) up to 1500 (diamond). Table 6.6
gives the Mohs scale with respect to the absolute hardness.
Common metals have Mohs hardness between 2 and 5 (Al: 2.75; Cu: 3.0;
iron: 4–5) so that with respect to wearing of polymer processing equip-
ment, only soft minerals close to talc and calcite are of interest as fillers.
Ground calcium carbonate, mica (Mohs hardness in the 2–2.5 range) and
clays (Mohs hardness about 2) give moderate wearing problems provid-
ing they do not contain hard impurities (e.g., sand or quartz particles).
Synthetic white fillers (e.g., precipitated calcium carbonate) have advan-
tages in this respect. Wollastonite, a form of calcium silicate, is a special
case, with a Mohs hardness of 4.8 but with needle-like particles that make
it interesting in certain applications. It must be noted however that if the
abrasivity of filler particles depends first on their hardness, their size and
266 Filled Polymers

Table 6.6
Mohs Hardness Scale
Mineral Chemical Composition Mohs Hardness Absolute Hardness
Talc (Mg3Si4O10(OH)2) 1 1
Gypsum (CaSO4·2H2O) 2 2
Calcite (CaCO3) 3 9
Fluorite (CaF2) 4 21
Apatite (Ca5(PO4)3(-OH,-Cl,-F) 5 48
Orthoclase (KAlSi3O8) 6 72
Quartz (SiO2) 7 100
Topaz (Al2SiO4(-OH,-F)2) 8 200
Corundum (Al2O3) 9 400
Diamond C 10 1500

shape are important factors as well. Indeed, particles with sharp edges, e.g.,
flakes, scales, or rod-shapes particles are more abrasive than smooth and
round particles, and molten polymer systems made with large particles
are generally more abrasive than with smaller ones. With respect to the
Mohs hardness of mineral glass (i.e.: 5.5) short glass fibers filled polymers
are likely to give the severest wearing problems, somewhat compensated
for however by using suitable surfactants and other additives. Other fac-
tors are important, such as the coefficient of friction, surface treatment and
energy (of both the filler particles and the metal of the processing equip-
ments), all are somewhat controllable, but the purity of the filler remains a
critical one since the most common contaminant in extracted minerals is
the highly abrasive sand (i.e., quartz).

6.2.2 Typical White Filler Effects and the Concept of Maximum


Volume Fraction
It is quite a trivial statement that, owing to their viscoelastic character, the
rheological and mechanical properties of polymers differ from those of other
materials, and it is quite a common observation that the viscoelastic charac-
ter of polymers is modified through the addition of foreign materials, e.g.,
mineral particles. The modification of the polymer viscoelastic behavior
results from (at least) three type of parameters:

• The filler volume fraction effect


• The shape and size of the particles
• Interactions between particles and/or between particles and polymer

No overall understanding exists for all possible combinations of those


effects in any given polymer; moreover effects or benefits obtained with
Polymers and White Fillers 267

a given filler-polymer system cannot generally be extended to another


­system. The preparation mode of the material deeply affects the quality of
dispersion of a given mineral in a given polymer and thermo-mechanical
degradation occurs rapidly if excessively energetic mixing processes are
used.
There is however an aspect which is qualitatively common to all
­filler–thermoplastic systems: the linear viscoelastic behavior exhibited
by most pure polymers at sufficiently low strain or low rate of deforma-
tion ­disappear above a sufficient filler level. For instance, the so-called
Newtonian plateau on the shear viscosity function is no longer observed,
the dynamic modulus is strongly strain dependent and the terminal region
in the elastic modulus function disappears and is replaced by a low fre-
quency plateau. As we have seen, such typical effects are also observed
with filled rubber compounds.
Let us consider the progressive disappearance, as the filler content increases,
of the pseudo-Newtonian region in the shear viscosity function. In extreme
cases, an (apparent) yield stress behavior occurs whose origin is assigned
to interactions between particles, which of course superimpose to hydrody-
namics effects. Particles interactions are nonhydrodynamic by nature and
depend on many factors, for instance the electric (ionic) and chemical prop-
erties of the particles, the surface area and properties of the particles and
the associated effects with the polymer matrix, and the presence of a sur-
factant (so-called compatibilizer) layer, if any. Yield stress behavior, either
apparent or real, is a low deformation rate phenomenon and, at higher defor-
mation rates, the nonhydrodynamic forces on the particles are dominated
by the stress field in the polymer matrix, which remains of course affected
by interparticles hydrodynamics, depending on the rheological properties
of the polymer and the arrangement of particles along flow lines. Different
arrangements of particles are expected in different types of flow and it is well
known that extensional flow fields induce an organization of particles that is
radically different from what is obtained in shear flow fields. Therefore par-
ticles arrangements, or in more general terms the flow induced anisotropic
distribution of particles, is strongly depending on the actual flow geometry.
For instance, in steady shear conditions, flow induced anisotropy in a paral-
lel disk rheometer is differing from the arrangement of particles in a cone-
plate system. All the more when extensional and shear flows are combined,
as in converging flow for instance, quite common in processing equipments
and techniques.
A common explanation for such phenomena is that the gradient of the
deformation rate tensor affects the flow induced anisotropy and hence the
stress. It follows obviously that filled (polymer) systems cannot be “sim-
ple” fluids since their behavior, by virtue of their true nature, violates the
principle of “local action,” which states that the stress in a fluid element is
determined by the deformation history of that fluid element and is indepen-
dent of the history of neighboring elements. This is the main reason for the
268 Filled Polymers

limited applicability of continuum mechanics to filled polymer systems; a


limit that can however be somewhat circumvented through micromechanic
approaches.
Let us consider a filled polymer system in a simple shear flow situation.
If indeed nonhydrodynamic interparticles effects dominate the shear flow
behavior at low shear rate and vanish at higher shear rates, leaving only
hydrodynamic effects, then one could consider that the actual viscosity
function of the filled material results from the mere superimposition of
two contributions, both expressed with the same convenient mathematical
model. Such models must explicitly consider a low shear viscosity plateau
and a high shear thinning behavior. The well-known Carreau–Yasuda model
meets such requisites and, as illustrated in Figure 6.12, the steady shear flow
behavior could be expressed through the following equation:54

(n1 − 1)/a1 ( n2 − 1)/a2


ηcpd ( γ ) = η0 ,1 1 + ( λ 1 γ ) 1  + η0 ,2 1 + ( λ 2 γ ) 2 
a a
(6.5)
   

The first term of the right member of this equation expresses the non-
hydrodynamic interparticles effects through parameters η0,1, λ1, a1 and n1;
the second term of the right member of this equation expresses the particles

Low shear
105 plateau η0,1
Apparent
yielding region
Shear viscosity, Pa.s

104 Intermediate plateau


ηc
η0,2
Shear
thinning
103 region

1 1 1
λ1 λc λ2

102
10–4 10–3 10–2 10–1 100 101 102 103
Shear rate, s–1

Figure 6.12
Modeling the shear viscosity function of filled polymer systems by combining two
Carreau–Yasuda equations; the curve was calculated with the following model parameters:
η0,1 = 8 × 104 Pa.s; λ1 = 500 s; a1 = 1.9; n1 = 0.4; η0,2 = 3 × 103 Pa.s; λ2 = 0.1 s; a2 = 3; n2 = 0.33.
Polymers and White Fillers 269

hydrodynamic effects through parameters η0,2, λ2, a2 and n2. The physical
meaning of parameters η0,1, λ1, η0,2, and λ2 is explicit from Figure 6.12. The
flow indices n1 and n2 monitor the shear thinning behavior in the low and
high shear regions respectively, and parameters a1 and a2 affect the curva-
ture of the viscosity function in the two transition regions.
Such a model allows thus four regions to be distinguished, with respect to
three critical shear rates:

1
• A low shear rate region, when γ ≤
λ1
1/a1
 a1 
1  1   η0 ,2  n1 − 1 
• An apparent yielding region, when λ ≤ γ ≤ λ   η  −1
1 

1

0 ,1 

1/a1
 a1 
1  η
 0 ,2  n1 − 1 1
− 1 ≤ γ ≤
• An intermediate plateau, when λ   η0 ,1   λ2
1  

1 
• A high shear thinning region, when ≤γ
λ2

Parameters λ1 and λ2 are characteristic times and, as clearly seen when con-
sidering the values of the parameters used to calculate the curve in Figure
6.12, the nonhydrodynamic interparticles effects (i.e., clustering) operate in a
time range that is several decades larger that the time range for polymer flow
processes. Available experimental data on filled polymers hardly meet such
a model in all its aspects, essentially because there is no technique to readily
capture the very low shear viscous behavior of very stiff systems. The author
has reported very low shear (down to the 10 –4 s–1 range) viscosity measure-
ments on filled rubber materials that do not show a yield stress limit but
suggest rather a low shear thinning region with an excessively high viscos-
ity. Similar observations are expected with filled thermoplastics. The very
low shear plateau, where nonhydrodynamic interparticles effects dominate
the viscous behavior, likely remains out of reach of experimental techniques
for the simple reason that in order to establish very low shear rate regimes,
one needs excessively long times. However, explicit yield stress σc data can
be extrapolated from low shear experiments and, if such data are available,
then Equation 6.5 reduces to:

σc a ( n − 1)/a
η( γ ) = + η0 1 + ( λ γ )  (6.6a)
γ  
270 Filled Polymers

If there is no intermediate plateau (corresponding to η0,2), one obviously


recovers the Herschel–Bulkley equation, i.e.:

σc
η( γ ) = + K γ n − 1 (6.6b)
γ

where the prefactor K has the meaning of the product η(1/λ) × λn–1. Despite
its mathematical simplicity, such a model offers large flexibility. For instance,
with the other parameters constant, how n1 and a1 values affect the shape of
the viscosity function is quite interesting. Indeed, either higher n1 or lower a1
somewhat dampen the intermediate plateau in such a manner that the shape
on the function appears very close to what can experimentally be observed
with filled polymers (see Appendix 6.2 for a numerical illustration).
The flow properties of concentrated suspensions in Newtonian fluids have
been studied for long and a few theoretical models have been compiled in
Table 5.12. However, not all conclusions of such studies can be extended to
filled polymer systems for at least two reasons: 1. only particles hydrody-
namic interactions are considered and 2. specific particle–matrix interactions
are not taken into consideration. Moreover, polymers are non-Newtonian
fluids and their viscoelastic character adds complexity. Certain aspects of
such studies are nevertheless interesting, namely particle clustering in shear
flow,55,56 and the associated problem of particles packing. For filled systems
that consist of even dispersions of non-interacting, rigid spheres of equal
diameter in a Newtonian matrix of viscosity η(0), it is indeed quite conve-
nient to consider that there is a maximum volume packing fraction Φm. For a
loose cubic packing, it is easy to establish that Φm = π/6 ≈ 0.5236 and for a close
hexagonal packing (also called face-centered cubic) of uniform spheres, one
has Φ m = π/3 2 ≈ 0.7405. Those two values can be considered at the lower
and upper bounds for the packing of uniform spheres. Other arrangements
give maximum packing fraction in between these two limits. For instance
the body-centered arrangement corresponds to Φ m = π 3/8 ≈ 0.6802 and for
a random close packing of uniform spheres, computer simulation yields Φm
≈ 0.64.* It is pretty obvious that with an adequate distribution of spheres’
diameters, the maximum packing distribution with the above ideal arrange-
ments is bound to increase. Vand57 was probably the first to hypothesize that
in concentrated suspensions, particles may cluster and that the suspending
liquid in the neighborhood of contact points between particles is effectively
immobilized and therefore contributes to an “effective” volume fraction

* The packing of objects (spheres, ellipsoids, marbles, etc) in a finite volume is a problem of con-
siderable interest, approached by mathematicians and physicists for centuries. It appeared
recently that the packing of spheres is apparently the exception rather than the rule, and that
as soon as the shape of objects becomes nonspherical, the packing efficiency increases by a
surprisingly large amount; see, for instance, D.A. Weitz. Packing in the spheres. Science, 303
(5660) 968–969, 2004.
Polymers and White Fillers 271

of the particles, larger that their true fraction. For clusters made of a small
­number of uniform sphere, simple geometrical arguments allow to estimate
the quantity of immobilized liquid. For larger cluster of spheres, it is conve-
nient to consider a “shape factor,” defined as the ratio of the actual volume
of the cluster of i particles to the overall volume of the i particles. As the
number of particles increases, such a shape factor becomes essentially con-
stant and depends only of the kind of packing. For model arrangements of
uniform spheres, it is easy to demonstrate that the shape factor is the reverse
of the maximum packing fraction, for instance equal to 1.9098, 1.4702, and
1.3505 for cubic, body-centered, and hexagonal packing, respectively.
For such ideal systems as suspensions of spheres of equal diameter, many
equations, either theoretical or empirical have been proposed for the relative
viscosity as a function of the filler volume fraction. Such a subject is obvi-
ously of tremendous importance in many fields. A thorough discussion of
suspensions of rigid particles in Newtonian fluids was made by Jeffrey and
Acrivos58 and some models available up to 1985 were discussed in detail
by Metzner.59 We will consider below only the most referred equations that
explicitly consider the maximum packing fraction. One of the oldest pro-
posal was likely made by Eilers60 in order to model the behavior of highly
viscous suspensions, i.e.:*

2
 5 
η( Φ )  1+ Φ 
= ηr ( Φ ) =  2  (6.7)
η( 0)  1− Φ 
 Φ m 

Obviously, owing to its simple quadratic form, the Eilers model is not
asymptotic to the Einstein’s one and therefore yields excessive values at low
volume fraction Φ. An often-quoted model, better in this respect, is the one
developed by Mooney for concentrated suspensions of uniform rigid spheres,
by considering only first-order interactions between spheres of equal diame-
ter, essentially a crowding effect. For very low spheres fraction, the Einstein’s
formula obviously applies and, by considering that a densely packed spheres
system would exhibit an infinite viscosity, Mooney established the following
equation:61

 
η( Φ ) 5 Φ 
= ηr ( Φ ) = exp   (6.8)
η( 0)  2 1− Φ 
 Φ m 

* Note that all the equations reproduced from literature are rewritten with respect to the for-
malism used throughout this book.
272 Filled Polymers

where Φm is the maximum volume packing fraction. Later, Krieger and


Dougherty modified the Mooney functional analysis to obtained an equa-
tion that fitted well experimental results on polymer latex. Their equation,
frequently quoted for suspensions of rigid particles, is:62

η( Φ )
− Φm B
 Φ 
= ηr ( Φ ) =  1 −  (6.9a)
η( 0)  Φ m

where B is a so-called “intrinsic viscosity” that should in fact match the


value 5/2 of the Einstein equation, as explicitly considered by Ball and
Richmond,63, i.e.:

5
η( Φ )
− Φm
 Φ 
= ηr ( Φ ) =  1 −
2
 (6.9b)
η( 0)  Φ m

Equations 6.9a or b are generally referred as the Krieger–Dougherty equa-


tion. A similar but simpler equation was proposed by Kitano et al.,64 i.e.:

η( Φ )
−2
 Φ 
= ηr ( Φ ) =  1 −  (6.9c)
η( 0)  Φm 

With respect to their intensive observations on uniform spheres in sim-


ple shear flow, Graham et al.65 paid attention to clustering of particles sus-
pended in Newtonian fluids and derived the following equation:

−2.5
η( Φ )    1− Φ    Φ m − Φ    
2

= ηr ( Φ ) = 1 − Φ 1 +  m
  1 −  Φ     (6.10)
η( 0)    Φ m   m   

With respect to the mathematical form of the well-known Guth, Gold, and
Simha equation, it is also interesting to mention a model previously devel-
oped by Graham:66

η( Φ ) 5 9 1 
= ηr ( Φ ) = 1 + Φ +  
η( 0) 2 4  A ( 1 + 0 . 5 Φ ) ( 1 + A )2 
  (6.11)
 Φ  Φ 
A = 2  1 − 3  Φ m 
3
  Φ m 

Figure 6.13 shows a comparison of these models, which in fact can be


divided in two groups: 1. models predicting a sharp variation of the relative
50 20
3 3

40 2 2
15
Polymers and White Fillers

30 1 1
0 0.1 0.2 0 0.1 0.2
10 Graham et al.
Eilers Mooney Krieger–Dougherty
20

Relative viscosity
Relative viscosity
5 Guth, Gold, Simha
Kitano
10 Graham
Guth, Gold, Simha
Einstein
Einstein
0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
Volume fraction Volume fraction

Figure 6.13
Comparing model equations for the relative viscosity of suspensions of uniform spheres.
273
274 Filled Polymers

viscosity upon increasing filler volume fraction, 2. Models predicting a mild


variation upon higher Φ. As can be seen, the former group consists of the
Eiler and Mooney equations and give predictions which are above the Guth,
Gold and Simha model (that does not consider a maximum packing frac-
tion); the later group is the Krieger–Dougherty, the Graham and the Graham
et al. equations, which are generally below the Guth, Gold and Simha model,
except the former at high volume fractions. Note that in calculating the
curves, the maximum hexagonal packing fraction for uniform spheres (i.e.,
0.74) was used. Using lower packing fraction somewhat modifies the shape
of the curves: the larger Φm, the larger the volume fraction range before rela-
tively viscosity goes to infinity for the Mooney and the Graham et al. models,
the lower Φm, the steeper the viscosity variation for the Krieger–Dougherty
and the Graham models (see Appendix 6.3).
If the filler has some trend to cluster in aggregates, one could somewhat
further develop the Krieger–Dougherty type of equations by considering an
effective enhanced volume fraction. Such an effective volume fraction could
then be related to both the overall size of the cluster D and the size of the
single particles d with respect to fractal considerations. Equation 6.8a would
then be rewritten as:

− Φm B
η( Φ )  Φ  D 
C
= ηr ( Φ ) =  1 −    (6.12)
η( 0)  Φ m  d  

where C is the so-called connectivity exponent from the fractal theory that
could for instance be taken as ≈ 1.8 (see Chapter 4, Section 4.1.4)
It is worth noting that the Krieger–Dougherty equation (Equation 6.8a) can
be expended in polynomial form to yield (see Appendix 6.5):

η( Φ ) ( B Φ m + 1) Φ 2 + B ( B Φ m + 1)( B Φ m + 2 ) Φ 3
= 1+ B Φ + B
η( 0) 2! Φm 3! Φ 2m

+B
( B Φ m + 1)( B Φ m + 2 )( B Φ m + 3) Φ 4 +  (6.13a)
4! Φ 3m

or, in an abridged form:

a−1

η( Φ )
n ∏(B Φ m + i)

η( 0)
= ηr ( Φ ) = 1 + B Φ + ∑B
a=2
i=1

a! Φ ma− 1
Φa (6.13b)
Polymers and White Fillers 275

If the series goes to infinity, Equation 6.13a and b match exactly the
Krieger–Dougherty equation and if B is taken equal to 2.5, when reduc-
ing the polynomial to the first two terms one obtains of course the
Mooney equation. This means that, in a very simple manner, expanding
the Krieger–Dougherty type of equation yields a polynomial function
that appears to somewhat take into account interactions between parti-
cles in addition to simple hydrodynamic effects. In agreement with the
theoretical reasonning by Einstein and Guld, Gold and Simha in deriv-
ing their equations, the three first terms of the polynomial account for
simple hydrodynamic effects and further terms account for interparticle
interactions, with the packing mode and its maximum packing fraction
playing the key roles. It is quite interesting to note that with B = 2.5 and
Φm = 0.74 (hexagonal packing of uniform spheres), the third and the fourth
terms in Equation 6.13b are respectively, 4.814 and 8.349, i.e., values fram-
ing the Φ2 multiplying factor in the relationship developed by Batchelor
for a suspension of spherical particles for which Brownian motion is
an issue,67 i.e.: ηr = (1 + 2.5 Φ + 6.2 Φ2) When compared with experimental
data, the Krieger–Dougherty is generally found to overpredict the filler
fraction effect, particularly when one approaches the a priori considered
maximum packing fraction. In the author’s opinion, the weakness of
the Krieger–Dougherty equation is that, with respect to its mathemati-
cal form and the logical limiting variation as Φ → 0, only the maximum
packing fraction is an adjustable parameter. This may lead of course to
quite unrealistic fitted Φm values. The polynomial equation, Equation
6.13, is quite attractive in this respect because, while keeping the a priori
considered values for B (i.e., B = 2.5) and Φm (i.e., with respect for instance
to information about the average shape of filler particles), the number of
terms of the polynomial allows to easily meet a large variety of experi-
mental observations, as illustrated in Figure 6.14. The number of needed
polynomial terms can then be interpreted as an indication of the extent of
interparticle interactions.
It is clear that the above models are oversimplified with respect to the
known complexity of the rheological behavior of suspensions and (obvi-
ously) of filled polymers. Their attractiveness is however their mathemati-
cal simplicity, whilst to consider that the volume fraction of the dispersed
particles is the only variable for the rheological properties is surely incor-
rect, notwithstanding the temperature, the mode of flow, the stress and the
rate of deformation that could however, in a first approximation, be consid-
ered as independent variables. In such a case, the shape, the average size
and the size distribution of the particles are obviously very influential fac-
tors, whose first effect will be to modify the maximum packing fraction.
For simple geometrical particle shapes (spheres, rods), Φm can be calculated
providing the arrangement is either geometrically defined or considered at
random (Table 6.7).
276 Filled Polymers

Uniform spheres, cubic packing


50 30
Φm = 0.74 Φm = 0.524

40 B = 0.25 B = 0.25
Relative viscosity

20

Relative viscosity
30

20
10

10

0 0
0 0.25 0.5 0.75 0 0.2 0.4 0.6
Filler volume fraction Filler volume fraction
Polynomial, 6 terms Polynomial, 3 terms
Polynomial, 12 terms Polynomial, 6 terms
Krieger–Dougherty Krieger–Dougherty
Guth & Gold

Uniform spheres, random packing Prolate ellipsoids, random packing


30 30

Φm = 0.621 Φm = 0.74
B = 0.25 B = 0.25
20 20
Relative viscosity

Relative viscosity

10 10

0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
Filler volume fraction Filler volume fraction
Polynomial, 3 terms Polynomial, 3 terms
Polynomial, 6 terms Polynomial, 6 terms
Krieger–Dougherty Krieger–Dougherty

Figure 6.14
Polynomial vs. Krieger–Dougherty equations for relative viscosity variations with respect to
maximum particle packing fraction.
Polymers and White Fillers 277

Table 6.7
Typical Maximum Volume Packing Fraction for Simple Geometries
Particles Shape Packing/Arrangement Fm
Spheres Cubic 0.524
all with the same diameter At random 0.601–0.640
Hexagonal 0.741
Spheres (bimodal) At random, D2/D1 = 3.8 0.68
two diameters D1, D2 At random, D2/D1 = 4.7 0.81
Rods At random 1
All with the same diameter 1.38 + 0.0376 f 1.4

f = shape factor f = 2 = > Φm = 0.676


f = 10 = > Φm = 0.430
f = 100 = > Φm = 0.040
Perfectly aligned π/(2 3 ) ≈ 0.907
Ellipsoids a,b,c [a: long axis] At random 0.68–0.71
Prolate ellipsoids [b >> c] At random 0.74

Table 6.7 reveals that aligned rods and randomly dispersed ellipsoids* can
pack more densely than spheres and the higher Φm for bimodal spheres is a
clear indication that particle size distribution is an important issue. Indeed,
it has long been observed that a dispersion of uniform spheres (i.e., same
diameter) has a higher viscosity than a suspension of spheres with different
diameters, while the volume fraction is kept constant.68,69 This would sug-
gest that a simple volume fraction Φ is insufficient to describe a polydisperse
suspension and that a distributive function Φ(d) would be more convenient,
with d a representative dimension of the particle. Such a distributive func-
tion is likely very difficult to assess and an alternative approach would be to
consider that the (experimentally determined) maximum packing fraction is
a typical information for a given filler. Indeed, good fitting of ηr (Φ) data for
filled polymers are in certain cases obtained if the packing fraction is consid-
ered as an adjustable parameter, as we will see later.
Equations 6.6 through 6.13 above (and many others in the literature)
are a priori valid only if the suspending medium is a Newtonian fluid,
i.e., whose viscosity is not affected by the rate of deformation. For non-
Newtonian suspending media, one of the simplest approach consists in
first accepting the above theoretical views about the role of the particles

* Note that quite complex simulation algorithms are needed to estimate the maximum packing
fraction of ellipsoids; for details, see A. Donev, I. Cisse, D. Sachs, E.A. Variano, F.H. Stillinger,
R. Connelly, S. Torquato, P.M. Chaikin. Improving the density of jammed disordered pack-
ings using ellipsoids. Science, 303 (5660), 990–993, 2004.
278 Filled Polymers

and second in introducing a multiplicative term to express the shear rate


dependence, e.g.70

−2 n− 1
 Φ  1 + ( λ cpd γ )2 
ηr ( Φ, γ ) = η0  1 − 2 (6.14)
 Φ m   

where η0 is the zero-shear viscosity of the polymer matrix and


λcpd = λ0 (1–(Φ/ Φm))–2 is a characteristic time of the compound, in fact the
characteristic time of the suspending medium multiplied by a function of
the filler fraction. In the term expressing the shear rate dependence, one
recognizes the Carreau equation for the shear viscosity function of poly-
mer melts. As such Equation 6.14 is relatively limited because it predicts
simple vertical shifting of the shear viscosity function of the polymer
matrix with increasing filler content. Experimental observation with filled
thermoplastics shows that this is not the case and that higher filler con-
tent not only shifts vertically the shear viscosity function but also modi-
fies its shape. Equation 6.14 lacks therefore flexibility and obvious changes
are first to introduce an explicit yield stress σc and second to refer to the
­Carreau–Yasuda model, i.e.

−2
σc  Φ  1 + ( λ cpd γ )a 
( n − 1)/a
ηr ( Φ, γ ) = + η0  1 − (6.15)
γ  Φ m   

As demonstrated in Figure 6.15 (see also Appendix 6.4), Equation 6.15 has
the capability to meet all the observed singularities that fillers impart to the
shear viscosity function. It is worth noting that to consider that a filled poly-
mer compound has a characteristic time that depends on the polymer, on
the filler fraction and its maximum packing fraction implies that the rheo-
logical (and also mechanical) properties of the system are strongly related
to a modification of the viscoelastic properties of the polymer matrix itself,
namely the spectrum of relaxation times, since by nature filler particles are
infinitely rigid when compared to the polymer. Certain experimental results
strongly support this point.71
In what filled molten polymers are concerned, it is worth underlining how-
ever that the maximum packing fraction remains essentially a theoretical
limit whose meaning must be somewhat changed. Indeed if, as it is frequently
the case, strong interactions between the polymer and the filler occur, or are
purposely initiated, the possibility of a full kinetic aggregation of particles
is obviously reduced, as least because of the high viscosity of the matrix.
But strong polymer–filler interactions promote also a reduced mobility of
polymer segments in the vicinity of particles’ surface. It follows that complex
polymer–filler clusters form, in which particles are bounded by polymer seg-
ments in a pseudo-glassy state. The actual and effective maximum packing
106 106
Φ = 0.1 η0 = 3 kPa.s λ0 = 0.01 η0 = 3 kPa.s
Φ = 0.3 λ0 = 0.01 s λ0 = 0.05 Φ = 0.2
105 105
Φ = 0.5 σc = 200 Pa λ0 = 0.60 σc = 200 Pa
n = 0.3 n = 0.3
104 a=2 104 a=2

Shear viscosity, Pa.s


Shear viscosity, Pa.s
103 103
Polymers and White Fillers

102 10 2
10–3 10–1 102 103 10–3 10–1 102 103
Shear rate, s–1 Shear rate, s–1

106 106
a = 0.5 η0 = 3 kPa.s σc = 100 η0 = 3 kPa.s
a = 1.0 λ0 = 0.01 s σc = 250 Φ = 0.3
105 105 λ0 = 0.01 s
a = 2.0 σc = 200 Pa σc = 500
n = 0.3 n = 0.3
104 a = 0.3 104 a=2

103 103

Shear viscosity, Pa.s


Shear viscosity, Pa.s
102 102
10–3 10–1 102 103 10–3 10–1 102 103
Shear rate, s–1 Shear rate, s–1

Figure 6.15
Capabilities of Equation 6.14 in meeting the typical features of the shear viscosity function of filled polymer systems; fixed parameters: η0 = 3 kPa.s,
n = 0.3, Φm = 0.74; variable parameters: Φ, λ0, a, σc.
279
280 Filled Polymers

fraction is consequently lower than the theoretical packing limit of the par-
ticles alone.

6.2.3  Thermoplastics and Calcium Carbonates


Calcium carbonate, either ground natural (GCC) or precipitated (PCC), is in
volume terms, the largest filler material for thermoplastics (over 70% of the
worldwide volume consumption), first in PVC but also in other polymers,
essentially polypropylene and polyamides. Pure CaCO3 is a relatively soft
material (Mohs hardness 3.0) so that wear of filled plastics processing equip-
ment is in principle limited, providing however there is no hard contami-
nant. Whatever the care and procedure in preparing GCC, it may contain
minute quantities of highly abrasive quartz (sand) depending on the deposit;
this is not the case for PCC. Depending on the particle size, GCC is cheaper
than PCC, or the reverse. Coarse GCC is obviously a cheap material but mill-
ing costs rise sharply as finer grades are produced until precipitated cal-
cium carbonate grades become the most economic to produce. A wide range
of particle sizes is available to suit many applications, with PCC the finest
range, and many grades are treated with hydrophobic agents, generally fatty
acids, in order to modify their surface or wetting properties and therefore,
to ease their dispersion in molten polymers, which are generally nonpolar.
Notable exceptions are polyamides for which uncoated (dry) CaCO3 grades
give the best properties. Both dry and wet coating processes are used by cal-
cium carbonate manufacturers. Dolomite is a calcium magnesium carbonate,
with widely spread deposits, that properly ground give grades that are used
in place of CaCO3 because it is slightly denser (2.85 vs. ~2.76 g/cm3), but it is
also slightly harder (Mohs hardness 3.5).
Worldwide, ready-to-be-processed CaCO3 filled polymers are available
from a number of compounders. Polypropylene compounds, in the 10–40%wt
range are available from numerous producers and grades up to 50%wt are
also locally offered for special applications. Table 6.8 gives the average prop-
erties of typical commercial PP–CaCO3 composites, as compiled from manu-
facturers’ data sheets (when available).
As can be seen, for a given loading in CaCO3, certain properties have aver-
aged values that suffer from such a large standard deviation that twice the
tabulated number could be considered as well. This obviously reflects the
diversity of polypropylene grades (see also the large standard deviations for
the based resins) used in preparing commercial composites, the presence
or the absence of coating agents and likely also the various compounding
procedures; such information are of course not disclosed by suppliers. As
expected the largest deviations are exhibited by fracture related proper-
ties, i.e., elongation at break and impact resistance. The melt flow index, the
ultimate tensile properties, the flexural yield strength tend to decrease with
increasing filler content, and the reverse is observed for the elasticity and
flexural moduli.
Table 6.8
Commercial PP–CaCO3 Composites; Average Suppliers’ Data

CaCO3 Content %wt 0 10 20 30 40 45 50


Calculated density* 0.96 1.02 1.07 1.11 1.13 1.15
0.087 0.112 0.124 0.135
Volume fraction Φ.** 0.032 0.061
0.090 0.127 0.147 0.157
Polymers and White Fillers

Property ASTM Unit


Specific gravity D792 g/cm3 0.90 ± 0.02 0.99 ± 0.07 1.04 ± 0.05 1.11 ± 0.07 1.26 ± 0.14 1.35 1.34 ± 0.02
Melt flow D1238 g/10 min 28 ± 18 13 ± 8 28 ± 37 15 ± 15 14 ± 15 16 –
Hardness, rockwell R 63 ± 37 76 ± 22 76 ± 24 78 ± 30 76 ± 29 – –
Hardness, shore D 60 ± 18 70 ± 4 69 ± 4 69 ± 4 72 ± 4 – –
Tensile strength, ultimate D638 MPa 26.6 ± 7 22.9 ± 9.8 23.3 ± 8.4 20.7 ± 7.2 20.6 ± 7.1 – –
Tensile strength, yield D638 MPa 26.2 ± 8.2 31.6 ± 18.8 23.7 ± 9.1 36.2 ± 30 22.1 ± 9 – 18 ± 2.6
Elongation at break D638 % 244 ± 35 229 ± 105 215 ± 117 203 ± 97 166 ± 35 – 143 ± 93
Elongation at yield D638 % 12 ± 11 33 ± 45 32 ± 46 18 ± 22 29 ± 41 – 5 ± 0
Tensile modulus D638 GPa 1.19 ± 0.35 1.50 ± 0.52 1.94 ± 0.73 2.19 ± 0.81 2.59 ± 0.81 – –
Flexural modulus D790 GPa 1.15 ± 0.42 1.77 ± 0.96 1.80 ± 0.79 2.37 ± 1.37 2.25 ± 1.51 2.5 2.30 ± 0.68
Flexural yield strength D790 MPa 30 ± 11 45 ± 31 38 ± 16 52 ± 40 36 ± 14 – –
Izod impact, unnotched D4812 J/cm 9.97 ± 8.56 10.02 ± 4.65 8.55 ± 4.53 10.62 ± 5.33 7.96 ± 4.45 – –
Gardner impact D256 J 12 ± 11 18 ± 20 18 ± 20 19 ± 20 16 ± 16 12 ± 0 16 ± 3

(Continued)
281
282

Table 6.8  (Continued)


CaCO3 Content %wt 0 10 20 30 40 45 50
Izod impact, notched J/cm 2.34 ± 3.83 2.17 ± 2.92 2.38 ± 3.41 1.94 ± 2.72 2.36 ± 3.6 – 0.78 ± 0.41
Izod Impact, notched ISO kJ/m2 2.2 ± 15 10 ± 7.4 9 ± 6.3 7.9 ± 5.5 2.1 ± 0.1 – –
HDT, at 0.46 MPa D648 °C 90 ± 25 92 ± 42 80 ± 47 100 ± 58 87 ± 53 – 101 ± 7
HDT, at 1.80 MPa D648 °C 52±6 72±37 56±34 71±29 61±33 – –
Vicat softening point °C 108±50 – 108±40 119±30 113±39 – –
* Theoretical densities were calculated with respect to the given weight percentage of CaCO3 and considering that the filler was coated with fatty
acid(s), 10% of its weight; the following densities were used: PP, 0.90; CaCO3, 2.76; fatty acid(s), 0.88 (average value for lauric, linoleic, myristic, oleic,
palmitic, and stearic acids).
** The calculated filler volume fraction depends on the exact composition of the composite (not given by suppliers) and on the density of the compound;
two values are given when the calculated and given densities are (too) different.
Filled Polymers
Polymers and White Fillers 283

In (isotactic) PP homopolymer, well dispersed fine CaCO3 particles


have been reported to give substantial increases in impact strength com-
pared to the unfilled polymer.72 But the toughening effect is strongly
depending on the hydrophobicity of the filler and, in this respect, fatty
acid treated grades have been found more interesting and the effect was
attributed to different interfacial phenomena imparted by the coating
layer.73 As shown in Table 6.8, the (average) impact resistance is indeed
somewhat improved when loadings below 20% are used, but the large
standard deviation suggests that for composites with certain grades of
virgin PP, the benefits might be substantially higher. Impact resistance is
however difficult to assess without a large experimental scatter and other
properties, somewhat related to toughness, show some significant trends
with increasing calcium carbonate content, as illustrated in Figure 6.16.
The tensile properties, either at yield or at break, steadily decrease with
higher filler level, but the flexural and elasticity moduli vary with CaCO3
level slightly above what is predicted by the Guth and Gold equation, i.e.,
G cpd = G polym(1 + 2.5Φ + 14.1Φ2). This is quite a remarkable observation since
it suggests that, in terms of flexural modulus, the effect of calcium carbon-
ate particles is slightly more than just hydrodynamic, with interactions
between the dispersed particles and the matrix playing a relatively minor
role. The left upper graph shows, in comparison with average commercial
data, flexural modulus results from a study with a 5.2 g/10 min MFI PP
and various levels of uncoated 4 μm CaCO3 particles (BET: = 18 m 2:g).74
The flexural modulus of the virgin resin is 0.567 GPa and using the Guld
and Gold equation, predicted flexural modulus values would be signifi-
cantly below the observed ones.
Generally speaking, the mechanical properties (stiffness, strength, and
toughness) of mineral filler-thermoplastic composites depend on three fac-
tors: filler particles size, particle/matrix adhesion, and filler loading. Particle/
matrix adhesion is especially important because strength depends on the
effective stress transfer between filler and matrix, and toughness/brittleness
is controlled by adhesion. The benefit in impact properties of CaCO3 -filled
PP have long been investigated and a direct correlation between the impact
strength and the matrix-to-filler adhesion has been recognized. Up to a cer-
tain filler level, the impact properties increase with increasing filler loading,
essentially because there is a weak adhesion of CaCO3 to the PP matrix, as
proved by published SEM micrographs of fracture surfaces.75 There are theo-
retical considerations that may help understanding this effect, for instance
the concept of critical interparticle distance76 and the associated preferential
orientation of crystal planes in the PP shell surrounding CaCO3 particles.77 If
the interparticle distance is lower than a certain critical value, then the com-
posite is expected to exhibit ductile rather than brittle fracture, and when the
interparticle distance is short enough, a preferential orientation of crystal
planes of the lowest shear resistance occurs between filler particles. Such a
crystallographic orientation is believed to lower the local plastic resistance,
284 Filled Polymers

Polypropylene + CaCO3
Modulus
4.0 Tensile properties

3.5 70

60 At yield
3.0

Tensile strength, at yield, MPa


Flexural modulus, GPa

2.5 50

2.0 40

1.5 30

1.0 20

0.5 10
Flexural
0 0
0 0.1 0.2 0.3 0 0.05 0.1 0.15 0.20
Filler volume fraction Filler volume fraction

3.5 40

35 At break
3.0
Tensile strength, ultimate, MPa

30
Elasticity modulus, GPa

2.5
25
2.0
20
1.5
15
1.0
10

0.5 5
Elasticity
0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.20
Filler volume fraction Filler volume fraction

Figure 6.16
Mechanical properties of commercial PP–CaCO3 composites; ° are averaged suppliers data; the
vertical bars indicate the standard deviations; shaded diamonds are data from one single man-
ufacturer; shaded triangles are data from B. Haworth, C.L. Raymond. Proceedings Eurofillers 97,
Manchester, UK, Sept. 8–11, 251–254, 1997; the curves in the left part have been calculated with
the Guth and Gold equation.

thus allowing larger local plastic deformation to occur and therefore an over-
all higher toughness of the composite.
In compiling data used to fill Table 6.8, no attention could be paid to the
quality of the calcium carbonate used in preparing the composites and par-
ticularly to the presence of a surface modifier, because this information is
Polymers and White Fillers 285

generally not disclosed by compounders. When using uncoated ­carbonate


fillers, impact resistance falls generally quickly with increasing filler level,
but when a coated filler is used, a very different behavior is observed:
notched impact strength remains higher up to 40%wt loading, i.e., 0.12 vol-
ume fraction. Coated CaCO3 (with a fatty acid, for instance) gives impact
resistance an order of magnitude greater than that of the unfilled poly-
mer. Easier dispersion and better processing properties of the composites
are obviously obtained with coated particles but at the expense of certain
mechanical properties. For instance, a PP composite made with 20% (vol-
ume) of 8% fatty acid coated CaCO3 exhibit a flexural modulus around
three times higher than the virgin polymer. If uncoated CaCO3 particles
are used, the composite modulus is more than four times higher. Using
4% of fatty acid gives nearly the same results. But using coated calcium
carbonate gives significant increases in impact resistance.74 Coating CaCO3
with a small amount of fatty acid (or a derivative) is generally considered
as beneficial, not only in terms of processing behavior, but also in term of
balance of mechanical properties, because enhanced impact resistance is
generally the main sought benefit. A common explanation for the benefit
of fatty acid (or other surface modifiers, e.g., organosilanes) is that a small
increase in the crystallization temperature is obtained, whilst maintain-
ing the same cristallinity because coating does not generally change the
nucleating power of the filler,78 already excellent if CaCO3 particles are fine
enough.79 Most of the benefit is obtained by using maximum 10% weight
surfactant (with respect to filler content). It must be noted however that
the coating effects may somewhat depend on the structure of the polymer.
Indeed, Supaphol and Harnsiri80 investigated recently the effects of CaCO3
particles of varying content, size, and type of surface modification on the
rheological and isothermal crystallization characteristics of syndiotactic
PP. They observed the usual effects on rheological properties, i.e., steady
shear and dynamic viscosity increasing with increasing content, decreas-
ing size, and surface coating of filler particles. The half-time of crystalliza-
tion was found to decrease with higher CaCO3 content and increases with
increasing particles size. When coating the particles with either stearic
acid or paraffin, their ability to effectively nucleate the polymer is signifi-
cantly reduced.
When coating calcium carbonate, the industrial approach consists so
far in using quite common and inexpensive agents, for instance paraf-
fins, fatty acids, and fatty acid derivatives. More expensive chemicals, e.g.,
silanes, titanates, seem rarely used in the industry, in contrast with the
abundant research literature on the benefits in using sophisticated surface
treating agents. Han et al.81 compared the effects of two silane coupling
agents (N-octyl triethoxy silane and -aminopropyl triethoxy silane), and of
isopropyl triisostearoyl titanate on rheological properties, processability,
and mechanical properties of 50% CaCO3-filed PP compounds. All cou-
pling agents give processing benefits (lower melt viscosity and higher melt
286 Filled Polymers

elasticity) but in quite different manners. For instance the two silanes affect
the melt elasticity in reverse manners. Mechanical properties of injection
molded specimens show the effects on the tensile strength and percent
elongation of the filled polypropylenes depend on the specific coupling
agent used. Leong et al.82 compared the effects of commercially available
neoalkoxy titanate, 3-aminopropyltriethoxysilane and stearic acid as cou-
pling agents in CaCO3 –PP composites. The silane and titanate treatments
dramatically increase the elongation at break for both the single-filler
and hybrid-filler composites, whereas stearic acid does not. A moderate
improvement in the impact strength of the composites is observed, par-
ticularly with the titanate product. Wang and Lee compared the benefits
of a liquid titanate coupling agent (isopropyl triisostearoyl titanate) and
stearic acid in CaCO3 –PP composites.83,84 A small amount of coupling
agents is found to give a drastic decrease in the surface energy of CaCO3
particles, and the magnitude of the effect depends on the type of coupling
agents. Completely covered particles exhibit low polarity surfaces. Infrared
analysis shows that stearic acid reacts extensively with the filler surface to
produce chemically bound organic salt compounds. No transesterification
reaction between the titanate agent and the filler is observed but IR data
suggest adsorption on the particle surface. Surface treated CaCO3 compos-
ites generally exhibit higher impact strength than untreated systems and
the titanate agent appears to me more effective than stearic acid. On a cost/
performance basis however, the low-cost stearic acid proves to be more
effective when dealing with impact properties.
In the author’s opinion, the exact mechanisms for the effects of the vari-
ous coating agents on the mechanical properties remain unclear however
whilst some (qualitative) relationships can be considered between the
modification of the rheological properties when coated CaCO3 particles are
used and the likely origin of the observed effects on mechanical proper-
ties. For instance, Price and Ansari85,86 used inverse gas chromatography
to characterize the surfaces of CaCO3 particles before and after treatment
with sodium polyacrylate or stearic acid, in order to explain the effects of
such coating agent on the mechanical properties of filled polypropylene. As
expected, they found that the surface treatment reduces the surface polar-
ity but also that modification with stearic acid produces nonpolar, low-
­energy surfaces. Some mechanical properties of the composites are found
to somewhat correlate with the surface energy modifications imparted
by the coating agents. Similar conclusions were reached by other authors
when considering the effects of titanate coupling agents on the rheologi-
cal properties of particulate-filled polyolefin melts.87 The main effect of the
coupling agent is a considerable reduction of the melt viscosity, at least in
the case of CaCO3 –PP systems, with some associated effects on mechanical
properties of injection molded specimens, for instance reduced modulus
and tensile strength, but increased elongation and impact strength of the
filled systems. The lowering of the surface energy of CaCO3 through stearic
Polymers and White Fillers 287

acid coating has been reported by other authors88,89 and the associated
benefits in ­processing behavior of PP–CaCO3 composites were assigned
to lower ­i nterfacial force between the filler surface and the resin matrix.
The benefits in using very fine particles for better impact resistance is well
established but, the smaller the particles, the more difficult the dispersion.
It is therefore, quite clear that an immediate benefit of coating agents, what-
ever their chemical nature, is to make CaCO3 particles easier to disperse by
decreasing their natural trend to remain clustered. As discussed by Richard
et al.90 the dispersion behavior of mineral fillers is strongly depending on
both the polar nature of the polymer and the surface properties of filler
particles. In a PVC matrix, CaCO3 particles rapidly disperse because there
are immediate strong acid-base interactions at polymer filler contacts. In
contrast, polypropylene (PP) is essentially a “Van der Waals-force material”
and fillers without pronounced acid or base surface characteristics tend
to disperse more rapidly and produce mechanically stronger compounds.
This would explain the beneficial aspect of coating agents on dispersion
mechanisms of CaCO3 in PP, with of course more decisive effects with very
fine particle materials, i.e., precipitated calcium carbonate.
It is worth underlining here a basic difference between filled elastomers
and filled thermoplastics. As we have seen, in the case of the former, the
normal loading for a reinforcing filler, e.g., carbon black or high structure
silica, is in the 50–60 phr range, that is in terms of volume fraction, largely
above the so-called percolation level, i.e., 0.12–0.13. In addition, a number
of interesting properties are due to the development of a soft “filler-rubber
network” embedded in the vulcanized matrix. Calcium carbonate filled
polypropylene grades are presented by manufacturers with respect to the
weight percentage of filler. In terms of mechanical (and rheological) prop-
erties, only volume fractions of ingredients are relevant and, accordingly,
the average data of Table 6.8 were plotted in Figure 6.16 with respect to
CaCO3 volume fraction. As can be seen, most of the commercial PP–CaCO3
grades have filler volume fraction below the percolation level. The obser-
vation that mechanical properties such as the flexural and the elasticity
moduli are slightly above the prediction of the Guth and Gold equation
indicates that, in addition to hydrodynamic effects between near spherical
particles, which do not interact much with each other (if dispersion is cor-
rectly made), there are other phenomena, whose origin is most likely the
boundary region between the particles and the matrix. In contrast with
carbon black filled rubber systems, the role of the interfacial region in filled
thermoplastics has so far received less attention, from both theoretical and
experimental points of views.
Although the modulus increases upon increasing filler content, the ten-
sile properties (strength at yield and at break in Figure 6.15; the correspond-
ing elongation data are given in Table 6.8) do not follow the same trend. As
illustrated above in the case of PP–CaCO3 systems, the tensile properties
decrease with increasing filler loading, due to stress concentration effects
288 Filled Polymers

(the so-called strain amplification effect, as seen in Chapter 5, Section 5.1.7),


but there are many factors that can affect these properties, for instance size
and geometry of the particles, interfacial adhesion, etc. For model systems
in which there is a good adhesion between the (spherical) particles and the
matrix, some authors91,92 reported that the elongation at yield (or at break)
should follow a simple relationship:

 
ε cpd = ε pol  1 −

3
3 2
π
3
Φ  ≈ ε pol ( 1 − 1.105

3
)
Φ (6.16)

where εcpd and εpol refer to the composite and the polymer respectively, and Φ
is the volume fraction of the filler. One notes that the prefactor 3 3 2/π
refers to a cubic close-pack array of uniform spheres. If there is poor adhe-
sion between the particles and the matrix, one would expect a more gradual
decrease in ultimate elongation with higher filler loading than that calcu-
lated with Equation 6.16. Figure 6.17 (left) shows that averaged data on com-
mercial PP–CaCO3 systems, despite the very large standard deviation, are not
far from the trend predicted by the equation. This would at least mean that,
whilst not exactly so, CaCO3 particles in a thermoplastic polymer roughly
behave as beads.
A rigid filler, finely dispersed in a rigid polymer, is normally expected to
decrease the impact strength of the composite, because such a property is
largely determined by dewetting and crazing phenomena. Interestingly, the
right part of Figure 6.17 shows the reverse: low volume fractions of calcium
carbonate somewhat improve the impact resistance, at least when notched
samples are used. A common interpretation of such an observation is that
nontouching, well dispersed CaCO3 particles would indeed promote craze
formation but would also impede crack growth, either through local higher
impact energy absorption effects or through crack propagation deviation. At
higher CaCO3 loading, this beneficial effect would be lost, as indeed shown
by Figure 6.17.
Approximately 80% of all the fillers used in PVC is calcium carbonate.
Titanium dioxide is second at around 12%, followed by calcined clay at about
5%. The remaining few percent is taken up by other materials, including
glass and talc. The performances of CaCO3 in PVC formulations strongly
depend on the particle size, the particle size distribution, the loading level
and the presence of so-called impact modifiers, for instance acrylic oligomers
or rubber-like materials, e.g., nitrile rubber NBR. Above a certain (average)
particle size (around 1 µm), PCC is at best a filler, below this size, ultrafine
PCC improves the impact resistance. Ultrafine PCC grades are commercially
available with particles in the 0.07 µm range.
Calcium carbonate, either ground or precipated, has been used as inex-
pensive filler and extender in flexible and rigid PVC formulations for more
than 30 years.93 As filler, both types of CaCO3 are nearly equivalent but
500
Polypropylene + CaCO3
20
400
18
16
Polymers and White Fillers

300 14
12
10
200
8

Elongation at break, %
6
100
4

(ISO) Izod impact, notched, KJ/m2


2
0 0
0.00 0.05 0.10 0.15 0.20 0 0.05 0.10 0.15 0.20
Filler volume fraction Filler volume fraction

Figure 6.17
Ultimate mechanical properties of commercial PP–CaCO3 composites; averaged suppliers data; the vertical bars indicate the standard deviations;
shaded diamonds are data from one single manufacturer; the curve in the left part has been calculated with Equation 6.16.
289
290 Filled Polymers

ultrafine (i.e., submicron particle size) PCC has revealed an interesting role
as impact modifier in rigid PVC.94,95 PCC particles are all about the same,
uniform in size, in contrast with GCC which has many large particles and
many very small particles. The narrow particle size distribution typi-
cal of PCC has many advantages, especially where high impact strength
is needed. As shown in Figure 6.18, impact strength improves when the
median particle size is below 1 µm. At low loading, the benefit of smaller
particle is relatively small, but above a critical loading, around 20 phr, the
benefit is particularly impressive. Note that CaCO3 contents were recalcu-
lated in terms of volume fraction using the formulation and specific gravity
data given in the caption of Figure 6.18. It is worth underlining that at 20
phr loading, the filler content is still far from the theoretical percolation
level (i.e., around 0.12).
As it has been reported that submicron PCC grades were enhancing the
gelation of PVC, it is likely that the origin of this beneficial effect on impact
resistance is related to improved dispersion not only of the filler particles but
also of the other compounding ingredients, namely the processing aid and
the lubricants. Indeed a better and more complete gelation would give a com-
posite with fewer defect sites and opportunities for crack propagation. Ductile
fracture, rather than brittle failure, is observed when using sufficient levels of
ultrafine precipitated calcium carbonate. The improvement in the impact per-
formance is obtained in rigid PVC compounds without the addition of any

1.8

1.6
PVC compounds
Izod impact strength, notched, kJ/m

1.4 CaCO content


phr Φ
1.2
10 0.046
1.0 15 0.068

0.8 20 0.088
Unfilled
0.6

0.4

0.2

0.0
0 1 2 3 4
Median particle size, µm

Figure 6.18
Effect of CaCO3 particle size on the impact resistance of rigid PVC (data from: Specialty Minerals
Inc., Easton, PA); Formulations: PVC: 100 (ρ = 1.39 g/cm3); stabilizer: 1.5 (1.05 g/cm3); acrylic
process aid: 1 (1.09 g/cm3); calcium stearate: 0.5 (1.04 g/cm3); fatty acid esters: 1.3 (0.96 g/cm3);
CaCO3: variable (2.71 g/cm3).
Polymers and White Fillers 291

organic impact modifier. But in the presence of either acrylic and rubber-type
modifiers, this improvement adds from the levels obtained with the modifier
alone. However, whereas polymer modified compounds display generally a
rapid fall in impact strength a low temperature, coated PCC modified sys-
tems still maintain a good impact resistance even at low temperatures.
Fine and ultrafine precipitated calcium carbonates are used in a num-
ber of rigid polyvinyl chloride applications such as window and door pro-
files, pipes and fittings, fencing and decking, increasingly replacing more
expensive impact modifiers. Besides providing improved impact resistance,
small particle PCCs give products with an excellent surface gloss and finish.
Coating PCC particles with fatty acid derivatives, e.g., fatty stearates, helps
the processing, namely during compounding through easier dispersion, and
gives lubrication effects during extrusion. It follows that, besides provid-
ing improved impact resistance, small particle PCCs give products with an
excellent surface gloss and finish.
In PVC cable fabrication, calcium carbonate is used at loadings of around
70 phr. The choice of calcium carbonate depends on the specifications for
cable. Higher quality cables benefit from the better mechanical properties
(tensile strength and elongation) and electrical properties (volume resistiv-
ity) offered by ultrafine coated grades. Cables with lower specification can
use uncoated PCC grades.

6.2.4  Thermoplastics and Talc


The term “talc” covers a wide range of products, but among its various forms,
mostly pure and lamellar grades are used in the plastic industry. Talc is a
crystalline form of magnesium silicate, whose uses in polymers is essentially
determined by its lamellar structure, essentially magnesium hydroxide sheets
between siloxane layers. Layers are bonded by weak Van der Waals forces
and therefore talc is a very soft material (Mohs hardness: 1) that, through
cleavage, gives high aspect ratio particles. It is the high aspect ratio that is
the most important property for its use in polymers. In thermoplastics, the
main usage of talc is in polypropylene where the high aspect ratio of par-
ticles gives significant increases in stiffness, but aspect ratio is rarely quoted
by talc suppliers. Depending on the deposit very pure and very white talcs
are available, with a great variety of average particle size and size distribu-
tion. Another variety of talc, generally referred as chlorite (a natural mixture
of magnesium silicate/hydroxide with iron and aluminium substitution) is
also used in polymers. No surface treatment has really been found of inter-
est when using talc as a filler for thermoplastics and consequently nearly all
marketed talcs are untreated. Talc is usually lamellar (platy), but the aspect
ratio may vary considerably depending on the deposit and the treatment
(grinding, sorting) of the filler.
There are many deposits and hence numerous producers of talc(s), but
around 80% of the world production come from seven countries: China,
292 Filled Polymers

South Korea, USA, India, Finland, France and Brazil. The former three
account for more than half of the world production (around 8.5 MioT in
2006). Talc is used in the manufacture of a myriad of products, with papers
and ceramics the largest markets worldwide (approximately 30% each).
Talc as a filler for thermoplastics (around 15% of the world consumption)
is in competition with fine grades of calcium carbonate (ground or precipi-
tated) and find applications in the production of automotive parts, house-
hold appliances, cables and engineering plastic parts. Talc has an average
growth of around 3% per year, essentially in line with the increasing use of
polypropylene, especially for the automobile market where lightweight and
recyclability are important issues nowadays. The development of very fine,
compacted, submicron grades offers the possibility to enhance the proper-
ties of plastic parts.
Talc is used to stiffen thermoplastics, mainly polypropylene but also poly-
ethylene and polyamide (nylon). In polypropylene compounds, talc is gener-
ally enhancing the following properties:

• Stiffness (elasticity modulus)


• Impact resistance (impact strength)
• Creep resistance
• Thermal conductivity
• Heat distortion temperature
• Nucleation
• Barrier properties
• Chemical resistance

Table 6.9 gives the average properties of typical commercial PP–talc com-
posites, as compiled from manufacturers’ data sheets (when available).
For a given filler loading, certain properties have averaged values that suf-
fer from a large standard deviation so that twice the tabulated number could
be considered as well. This obviously reflects the diversity of polypropylene
grades used in preparing commercial composites, the particular grade of
talc and possibly some additives not disclosed by the manufacturers. The
compounding procedure also plays a role. The largest deviations are exhib-
ited by fracture related properties, i.e., elongation at break and impact resis-
tance. The melt flow index and the elongations at break and at yield tend to
decrease with increasing filler content, but the reverse is observed for the
elasticity and flexural moduli.
The main reason for incorporating talc in polypropylene is to increase
the stiffness (modulus of elasticity). The degree of rigidity obtained
depends on the filling level, aspect ratio and fineness of the talc, with
high aspect ratio grades giving the largest enhancement. Figure 6.19
shows how the flexural and elasticity moduli vary with increasing talc
Table 6.9
Commercial PP–talc Composites; Average Suppliers’ Data
Talc Content %wt 0 10 20 30 40 50
Calculated density* 0.964 1.019 1.071 1.12 1.166
0.0318 0.0615 0.0895 0.1159 0.1408
Volume fraction Φ.**
0.0323 0.0707 0.0995 0.1304 0.1655
Polymers and White Fillers

Property ASTM Unit


Specific gravity D792 g/cm3 0.90 ± 0.02 0.98 ± 0.03 1.17 ± 0.25 1.19 ± 0.17 1.26 ± 0.06 1.37 ± 0.03
Melt flow D1238 g/10 min 28 ± 18 17 ± 12 17 ± 12 12 ± 6 14 ± 15 7 ± 6
Hardness, rockwell R 63 ± 37 82 ± 20 79 ± 24 75 ± 28 76 ± 31 90
Hardness, shore D 60 ± 18 70 ± 4 71 ± 9 73 ± 9 74 ± 10 –
Tensile strength, ultimate D638 MPa 26.6 ± 7.0 25.2 ± 11.1 24.6 ± 11.7 24.5 ± 10.6 18.9 ± 12.4 23.5 ± 2.1
Tensile strength, yield D638 MPa 26.2 ± 8.2 27.7 ± 10.1 27.1 ± 10.4 26 ± 9.5 35.3 ± 22.5 30.4 ± 5.8
Elongation at break D638 % 169 ± 83 31 ± 21 13 ± 14 13 ± 9 13 ± 7 3 ± 3
Elongation at yield D638 % 12 ± 11 12 ± 8 10 ± 7 9 ± 6 3 ± 2 3 ± 1
Tensile modulus D790 GPa 1.19 ± 0.35 1.74 ± 0.82 2.26 ± 1.13 2.49 ± 1 3.26 ± 1.4 3.75 ± 0.64
Flexural modulus D790 GPa 1.15 ± 0.42 1.56 ± 0.4 2.09 ± 0.83 2.18 ± 0.65 3.23 ± 0.5 4.14 ± 0.86
Flexural yield strength D790 MPa 30 ± 11 37 ± 16 38.7 ± 17.2 44.2 ± 14.7 43.8 ± 22.1 –
Izod impact, unnotched D4812 J/cm 9.97 ± 8.56 5.82 ± 3.62 3.91 ± 3.11 2.85 ± 2.24 2.37 ± 1.99 –
Charpy impact, notched J/cm2 – 0.45 ± 0.21 1.5 ± 0.07 0.22 ± 0.06 0.24 ± 0.14 –
Gardner impact D256 J 12 ± 11 2 ± 3 1 ± 0 3 ± 2 5 ± 7 2

(Continued)
293
294

Table 6.9  (Continued)


Talc Content %wt 0 10 20 30 40 50
Izod impact, notched J/cm 2.34 ± 3.83 3.65 ± 1.78 3.54 ± 2.22 2.9 ± 1.72 1.14 ± 0.83 0.27 ± 0.07
Izod impact, notched ISO kJ/m2 2.2 ± 15 6.5 ± 4.2 2 ± 0 8 ± 6.3 6 ± 5.3 –
Deflection temperature D648 °C 90 ± 25 137 ± 74 109 ± 30 104 ± 44 126 ± 19 140 ± 2
Deflection temperature D648 °C 52±6 72±63 86±38 85±46 93±44 –
Vicat softening point °C 108±50 104±12 110±42 113±72 116±40 158

* Theoretical densities were calculated with respect to the given weight percentage of talc and considering that the filler was uncoated; the following
densities were used: PP, 0.90; talc, 2.76.
** The calculated filler volume fraction depends on the exact composition of the composite (not given by suppliers) and on the density of the compound;
two values are given when the calculated and given densities are (too) different.
Filled Polymers
Polymers and White Fillers 295

Polypropylene + talc
6 6

5 5

Elasticity modulus, GPa


Flexural modulus, GPa

4 4

3 3

2 2

1 1

0 0
0 0.05 0.10 0.15 0.20 0 0.05 0.10 0.15 0.20
Filler volume fraction Filler volume fraction

Figure 6.19
Effect of talc volume fraction on the flexural and elasticity moduli of commercial filled poly-
propylene compounds; clear squares are average moduli data; gray shaded diamonds are data
from one selected manufacturer; the dotted curve was calculated with the Guth and Gold equa-
tion and Epolym = 1.152 GPa and 1.191 GPa for the flexural and elasticity moduli respectively; the
solid curve was calculated with the modified equation and an anisometry factor f = 2.1.

volume fraction. Talc particles are far to correspond to spheres and there-
fore the variation is larger than predicted with the Guth–Gold equation.
A simple manner to account for particles anisometry, while considering
that the filler effect is essentially hydrodynamic, consists in introducing
an anisometry factor f, i.e.

Ecpd = Epolym ( 1 + 2.5 × f × Φ + 14.1 × f 2 Φ 2 ) (6.17)

As shown in Figure 6.19, this remarkably simple approach suits well aver-
age moduli data on talc filled polypropylene compounds.
Compared to polymers, talc has a significantly higher thermal conductiv-
ity and therefore, heat transfers during processing are more efficient: talc
filled PP melts easier and cools faster and high production rates can be
obtained. Small quantities of fine talc particles act as a nucleating agent, thus
promoting the crystallization of polypropylene.96–99 For instance, exothermic
crystallization peaks, as measured by DCS, show that talc particles, whilst
having nearly twice as large average particle size are stronger nucleating
agents of PP than CaCO3 and kaolin.75 When the crystallinity of the poly-
mer matrix increases, the composite is expected to exhibit a higher modulus,
296 Filled Polymers

an increased tensile strength and a better dimensional stability, as indeed


observed. Talc is better than kaolin and CaCO3 in this respect. Crystallization
starts at a higher temperature in the presence of talc, compared to unfilled
PP. The impact strength is improved but this is primarily due to an increase
in the crystallization of the PP and not the mechanical properties of the talc
itself. There is also a change in modulus as a result of the change in crystal-
linity. Mineral fillers do not generally improve impact strength, but fine talc
is an exception in PP compounds that explain its use in manufacturing car
bumpers. In this case, 5–10% of fine talc give a benefit, which however van-
ishes at higher loadings, as indeed shown in Figure 6.20, drawn with average
Izod data (notched specimen, test at room temperature). Other impact tests
(average) data show a similar trend.
Talc has a beneficial effect on the deflection temperature of PP parts, but it
strongly depends on the aspect ratio of the particles, as easily understood via
micromechanics considerations. Mineral fillers impart substantial reduction
in creep of plastics and fine platy talc grades give the best performance, with
however some synergistic effects when additional different fillers (e.g., cal-
cium carbonate) are used, without or with surfactant additive.100,101 Thanks
to its lamellar structure, talc has excellent barrier properties, by reducing
for instance transmission rates for water vapor and oxygen, providing how-
ever that particles are properly oriented during the processing operations.
Lamellar talc particles are mostly orientated in films and help constraining

6
Izod impact, notched, J/cm

0
0 0.1 0.2
Filler volume fraction

Figure 6.20
Effect of talc volume fraction on the impact resistance of commercial filled polypropylene com-
pounds; clear squares are average moduli data; shaded diamonds are data from one selected
manufacturer.
Polymers and White Fillers 297

Table 6.10
Typical Industrial Applications of Talc in Thermoplastics
Typical Median
Loading Typical Particle
Application Role (%) Grade Size (µm)
Additive in
semicrystalline Nucleating agent 0.5 Very pure 1–2
polymers
Controlled size
Plastic films Antiblocking agent 0.5 2–3
distribution
LDPE/EVA agro-films
IR radiation retainer 5–10 Very pure 5
(greenhouses)
Domestic appliances
White reinforcing filler 30–40 Very white ≈ 10
(injection molded parts)
Household appliances Temperature
20 Very white 6–10
(PP) resistance
Garden furnitures (PP) Stiffness, rigidity 20 Very white 10
Stiffness, gas
Industrial PP foils/films 20 Lamellar 3–6
impermeability

Automotives Parts
Parts in engine Temperature
40 Lamellar, pure 10
compartment resistance
Dashboards Dimensional stability 15–20 Lamellar 3–5
Stiffness, impact
Bumpers, exterior parts 5–10 Fine, lamellar 2
strength

the water vapor and oxygen. Talc is water repellent and chemically inert and
therefore increases the chemical resistance of PP parts.
The above data explain why the automotive and domestic appliances
markets are the major ones for talc-filled PP. Talc is the preferred additive
for such applications, as it imparts high stiffness, which allows a reduc-
tion in wall thickness, at least below a certain loading. Food packaging
applications for talc-filled PP systems are due to the higher rigidity and
barrier properties obtained. Table 6.10 describes a few typical industrial
usages of talc.

6.2.5  Thermoplastics and Mica


Mica is considered as the third most important white filler for thermoplas-
tics but has some unique characteristics that sometimes give filled materi-
als exceptional properties not obtained with any other mineral. Mica is a
transparent, flexible and elastic phyllosilicate that can be ground to very
fine particles with a high-aspect ratio and high mechanical properties, e.g.,
tensile modulus ≈ 175 MPa, flexural modulus ≈ 220–270 MPa, compressive
298 Filled Polymers

modulus ≈ 190–285 MPa (as measured on sheets of muscovite). Mica is rela-


tively soft, with Mohs hardness of around 2.8–3.2 that helps in minimizing
wear in plastic processing equipment. Finely ground mica has a high degree
of whiteness (around 83%) and is strongly absorbing UV radiations, thus
increasing the service life and decorative value of end products for outdoor
applications. In semicrystalline thermoplastics, mica is believed to act as a
nucleating agent.
Polypropylene is likely the major polymer compounded with mica for a
number of applications, mainly in the automobile industry, sometimes in
association with other white fillers, calcium carbonate, talc or short glass
fibers, to fabricate a large variety of parts, e.g., instrument panels, heater
housing, rear-light retainers, ceiling fan blades, etc. Other applications
for mica-filled PP compounds are: blow molded containers and consumer
goods packaging, extruded fibers, filaments and films, pipes and sheets,
wires and cables), injection molded parts, e.g., furniture, house wares, lug-
gage, boxes, etc.
Compounded mica filled polypropylenes are available from a number
of manufacturers, with however a large offer from India (where are the
largest deposits of mica). Grades up to 40% are marketed and Table 6.11 is
a compilation of typical average properties, based on manufacturers’ data
sheets.
As can be seen, both the tensile and flexural properties (strength and
modulus) generally increase with higher mica content, with corresponding
decreases in elongation (at yield and at break). Mica has a variable and gener-
ally negative effect on impact strength. As expected, heat distortion tempera-
ture markedly increases with increasing filler content. Flexural and elasticity
moduli are worth special attention. Figure 6.21 compares data from Table
6.11 and from a selected compounder with the prediction of the Guth and
Gold equation, modified by the addition of an anisometry factor of 3.3. As
can be seen, the agreement is excellent which suggests that the prime effect
of mica particles is essentially hydrodynamic, with however the flaky shape
playing a key role.
The incorporation of mica generally reduces thermal expansion and helps
eliminating the nonuniform thermal shrinkage in injection molding; both
effects are obviously associated with the natural flake-like form of ground
mica. Like talc, mica has also excellent barrier properties, all the more if
the particles are properly oriented with respect to the trajectory of the
permeant.
Mica provides similar benefits in a wide range of thermoplastic and ther-
moset composites including polyolefins, polyamides and styrenics. It is also
reported that surface coated mica further increases tensile strength, flexural
strength and modulus, and heat deflection temperature. The automobile
industry is the main user of mica-filled composites, either with polypropyl-
ene or nylon as polymer matrix. Up to 40% mica loadings are used, some-
times in association with calcium carbonate, to produce various injection
Table 6.11
Commercial PP–mica Composites; Average Suppliers’ Data
Mica, %wt 0 10 20 25 30 40
Calculated density 0.97 1.026 1.053 1.079 1.128
0.0313 0.0606 0.0747 0.0883 0.1143
Volume fraction range
0.0309 0.0615 0.0706 0.0925 0.1307
Density g/cm3 0.90 ± 0.02 0.96 ± 0.01 1.04 ± 0.03 1.08 ± 0 1.13 ± 0.04 1.29 ± 0.11
Melt flow g/10 min 28 ± 18 8 ± 7 11 ± 9 – 6 ± 5 16 ± 5
Hardness, rockwell R 63 ± 37 65 ± 19 55 ± 22 97 43 ± 20 88 ± 23
Polymers and White Fillers

Hardness, shore D 60 ± 18 70 ± 1 74 ± 4 – 71 ± 1 73 ± 1


Tensile strength, ultimate MPa 26.6 ± 7.0 18.5 ± 7.7 24.8 ± 10.7 32.6 ± 3.7 17.2 ± 7.3 26.6 ± 8
Tensile strength, yield MPa 26.2 ± 8.2 28.4 ± 2 23.2 ± 10.2 34.2 ± 1.1 31.2 ± 7.1 38.3 ± 7.1
Elongation at break % 169 ± 83 19 ± 3 12 ± 5 15 ± 1 4 ± 3 10 ± 3
Elongation at yield % 12 ± 11 6 ± 1 5 ± 2 4 2 ± 1 4 ± 1
Tensile modulus GPa 1.19 ± 0.35 1.37 ± 0.38 2.1 ± 0.46 3.62 ± 0.68 2.86 ± 0.32 4.69 ± 1.11
Flexural modulus GPa 1.15 ± 0.42 1.56 ± 0.64 2.48 ± 0.68 3.52 ± 0.1 3.13 ± 0.94 5.24 ± 1.67
Flexural yield strength MPa 30 ± 11 28.8 ± 11 35.6 ± 12.7 48 38.2 ± 5.4 52.5 ± 9.1
Izod impact, unnotched J/cm 9.97 ± 8.56 – 3.88 ± 1.27 4.81 1.67 ± 0.64 3.47 ± 1.27
Gardner impact J 12 ± 11 1.6 ± 1.8 0.8 ± 0.7 – 0.6 ± 0.4 1.8 ± 0.7
Izod impact, notched J/cm 2.34 ± 3.83 0.58 ± 0.37 0.77 ± 0.81 0.39 ± 0.13 0.32 ± 0.18 0.73 ± 0.52
HDT, at 0.46 MPa °C 90 ± 25 87 ± 18 116 ± 19 124 106 ± 13 141 ± 10
HDT, at 1.8 MPa °C 52 ± 6 54 ± 7 68 ± 13 69 65 ± 11 102 ± 22
* Theoretical densities were calculated with respect to the given weight percentage of (uncoated) mica; the following densities were used: PP, 0.90;
mica, 2.82.
** The calculated filler volume fraction depends on the exact composition of the composite (not given by suppliers) and on the density of the compound;
two values are given as the calculated and given densities are different.
299
300 Filled Polymers

Polypropylene + mica
7 7

6 6

5 5

Elasticity modulus, GPa


Flexural modulus, GPa

4 4

3 3

2 2

1 1

0 0
0.00 0.05 0.10 0.15 0.00 0.05 0.10 0.15
Filler volume fraction Filler volume fraction

Figure 6.21
Effect of mica volume fraction on the flexural and elasticity moduli of commercial filled poly-
propylene compounds; clear squares are average moduli data; shaded diamonds are data
from one selected supplier; the dotted curve was calculated with the Guth and Gold equation
and Epolym = 1.152 GPa and 1.191 GPa for the flexural and elasticity moduli, respectively; the
solid curve was calculated with the modified Guth and Gold equation and an anisometry
factor f = 3.3.

molded PP parts, such as instrument panels, heater housing, ceiling fan


blades, tail-light retainers, door panels and seat backs in automobiles. Mica
is also used as a filler in fire-retardant polypropylene foams, for instance in
side components for underground swimming pools. Mica-filled nylons are
used to produce automobile parts such as opening panels, various parts for
the engine compartment and also headlight covers. In such applications,
mica generally provides higher tensile strength, heat deflection tempera-
ture and notched Izod impact than other mineral fillers, such as CaCO3 and
wollastonite.

6.2.6  Thermoplastics and Clay(s)


Kaolin (clay) is a widely used filler for thermoplastic polymers because
of its reinforcing effects on mechanical properties, such as stiffness and
strength, with however certain negative effects on impact resistance. Clays
are typically used at 20–50% wt in polymer composites, and commercially
available grades have aspect ratio between 4 and 12, and average particles
in the 0.5–2.3 µm range. Because kaolin has a platelike structure with a high
Polymers and White Fillers 301

aspect ratio, particle size and particle sizes distribution are very important
parameters in determining the mechanical properties of filled composites.
Kaolin improved the stiffness of PP102 and other polymers, and surface
treatment of particles can somewhat improve the mechanical properties of
the composites,103 likely because either the dispersion is easier during the
compounding operations or because a good adhesion is obtained at the
filler–matrix interface, or both. Clays are phyllosilicates and therefore, can
be delaminated through the use of suitable chemicals and/or mechanical
means. By splitting apart the platelet structure, which further increases the
available surface area more filler–polymer contacts are obtained at constant
loading, and therefore enhanced reinforced properties are obtained. It has
been shown that dramatic improvements in mechanical properties can be
achieved with very low levels (a few weight percent) of inorganic exfoliated
clays. Exfoliated clays have a thickness of around 1 nm and lateral dimen-
sions of ≈ 30 nm to several microns or larger. The large aspect ratios of exfo-
liated clays are thought to be the main reason for the enhanced mechanical
properties of so-called polymer nanocomposites. Considerable research
work has been published on polymer nanocomposites, with a few recent
reviews104,105 that are worth reading, but sizeable industrial applications
remain scarce. The subject is outside the scope of the present book and will
not be discussed further here.
Combining several (white) fillers sometimes allow to obtain composites
with a wider range of properties at acceptable costs. Calcium carbonate pro-
vides improved impact properties in polyolefins but relatively modest stiff-
ness. Talc, mica and clay, owing to their aspect ratio, give moderate stiffness
but with adverse effects on impact resistance. Short glass fibers provide a
high stiffness, somewhat depending on average fiber orientation, but bring
a considerable brittleness. Through a judicious and adequate combination
of the above filler materials, it is thus in principle possible to achieve a very
broad range of properties, but in a highly pragmatic manner. For instance,
so-called engineering plastics, e.g., PET, PBT, and ABS, can be filled with up
to 30% mica, in association with 20–25% short glass fibers to obtain improved
tensile, flexural and compressive strength, stiffness and to impart particu-
lar electrical properties. Except for a few very low demanding applications,
adding fillers to a thermoplastic polymer does not cheapen formulations, at
least on a volume basis because mixing and compounding costs ­generally
override savings in raw materials. But in many cases, the overall balance
of properties achieved with filled polymers is such that the composite may
compete with more expensive, high performance engineering materials.
Good examples are filled-reinforced PP to compete with ABS, and short
glass ­fiber-reinforced polyamides to compete with more expensive specialty
polymers. This likely explains the huge diversity of filled thermoplastics
available today, as a demonstration of continuously extended opportunities
for filled polymer composites.
302 Filled Polymers

References
1. M.P. Wagner. Reinforcing silicas and silicates. Rubb. Chem. Technol., 49, 703–774,
1976.
2. B.B. Boonstra, H. Cochrane, E.M. Dannenberg. Reinforcement of silicone rubber
by particulate silica. Rubb. Chem. Technol., 48, 558–576, 1975.
3. W. Lynch. Handbook of Silicone Rubber Fabrication, 2nd Ed. Van Nostrand Reinhold,
New York, NY, 1978. ISBN 978-0442249625.
4. S. Wolff. Chemical aspects of rubber reinforcement by fillers. Rubb. Chem.
Technol., 69, 325–346, 1996.
5. B. Freund, W. Niedermeier. Molekulare Deutung des Payne-Effektes und
Beeinflussung durch Füllstoffe. Kautsch. Gummi Kunstst., 51, 444–449, 1998.
6. S. Wolff, U. Görl, M.J. Wang. Silica-based tread compounds. Eur. Rubb. J., 16 (1),
16–19, 1994.
7. T. Kataoka, P.B. Zetterlund, B. Yamada. Effect of mixing sequence on the properties
of carbon black and silica filled rubber. Plast. Rubb. Comp., 32 (7), 291–296, 2003.
8. A. Voet, J.C. Porawski, J.B. Donnet. Reinforcement of elastmoers by silica. Rubb.
Chem. Technol., 50, 342–355, 1975.
9. A. Hunsche, U. Görl, A. Müller, M. Knaack, Th. Göbel. Investigations concern-
ing the reaction silica/organosilane and organosilane/polymer. Part 1. Reaction
mechanism and reaction model for silica/organosilane. Kautch. Gummi Kunstst.,
50, 881–889, 1997.
10. M.J. Wang, S. Wolff, J.B. Donnet. Filler-elastomer interactions. Part I: silica sur-
face energies and interactions with model compounds. Rubb. Chem. Technol., 64,
559–576, 1991.
11. U. Görl, A. Hunsche. Advanced investigations into the silica/silane reaction
system. ACS, Rubb. Div. Mtg., Louisville, KY, Oct. 8–11, 1996. Paper 76.
12. U. Görl, A. Hunsche, A. Mueller, H.G. Koban. Investigations into the silica/
silane reaction system. Rubb. Chem. Technol., 70, 608–623, 1997.
13. A. Hunsche, U. Görl, H.G. Koban, T. Lehman. Investigations concerning the
reaction silica/organosilane and organosilane/polymer. Part 2. Kinetic aspects
of the silica-organosilane reaction. Kautch. Gummi Kunstst., 51, 525–533, 1998.
14. P. Pal, S. De. Studies of polymer-filled interaction, network structure, physical
properties, and fracture of silica-and-clay-filled EPDM rubber in the presence of
a silane coupling agent. Rubb. Chem. Technol., 56, 737–773, 1983.
15. U. Görl, A. Parkhouse. Investigations concerning the reaction silica/organosi-
lane and organosilane/polymer. Part 3. Investigations using rubber compounds.
Kautch. Gummi Kunstst., 52, 493–500, 1999.
16. S. Wolff. Reinforcing and vulcanization effects of silane Si-69 in silica-filled com-
pounds. Kautsch. Gummi, Kunstst., 34, 280–284, 1981.
17. A. Blume, M. Janik, J.P. Gallas, F. Thibault-Starzyk, A. Vimont. Operando infra-
red study of the reaction of triethoxypropylsilane with silica. Kautch. Gummi
Kunstst., 61 (7–8), 359–362, 2008.
18. W.K. Dierkes. Economic mixing of silica-rubber compounds—Interaction
between the chemistry of the silica-silane reaction and the physics of mixing.
PhD Thesis, University of Twente, Enschede, The Netherlands, 2005. ISBN
90-365-2185-8.
Polymers and White Fillers 303

19. G. Nijman, J.L. Leblanc. Engineering performance and material viscoelastic


analyses alonga compounding line for silica based compoundspart 1: mixing
line performance analysis. DIK, 8th Fall Rubber Colloquium, Hannover, Germany,
Nov. 26–28, 2008.
20. R. Rauline, patent assigned to Michelin & Cie, Clermond-Ferrand, France.
Rubber compound and tires based on such a compound. Eur. Patent EP0501227,
Sept. 1992.
21. G. Heinrich, T.A. Vilgis. Why silica technology needs S-SBR in high performance
tires? Kautsch. Gummi Kunstst., 61, 368–376, 2008.
22. A.A Ward, A.A. Yehia, A. M. Bishai, F.F. Hanna, A.A. Mansour, B. Stoll, W. von
Soden, S. Herminghaus. Dynamic-mechanical properties of solution sturene
butadiene rubber loaded with silica. Kautsch. Gummi Kunstst., 61, 569–575,
2008.
23. G. Berrod, A. Vidal, E. Papirer, J. B. Donnet. Reinforcement of siloxane elastomers
by silica. Chemical interactions between an oligomer of poly(dimethylsiloxane)
and a fumed silica. J. Appl. Polym. Sci., 26, 833–845, 1981.
24. G. Berrod, A. Vidal, E. Papirer, J.B. Donnet. Reinforcement of siloxane elasto-
mers by silicas. Comparison between fumed and precipitated silicas in their
interactions with an oligomer of poly(dimethylsiloxane). J. Appl. Polym. Sci., 26,
1015–1025, 1981.
25. S. Ross, G. Nishioka. Monolayer studies of silica/polydimethylsiloxane disper-
sions. J. Colloid Interface Sci., 65, 216–224, 1978.
26. J.P. Cohen-Addad, N. Morel. NMR investigations into polydimethylsiloxane adsorp-
tion on silica aggregates. J. Physique III, 6 (2), 267–277, 1996.
27. J.P. Cohen-Addad. Silica-siloxane mixtures. Structure of the adsorbed layer:
chain length dependence. Polymer, 30, 1820–1823, 1989.
28. T. Cosgrove, M.J. Turner, D.R. Thomas. The adsorption of polydimethylsiloxane
onto silica from the melt. Polymer, 38, 3885–3892, 1997.
29. R. Al Akoum, B. Haidar, A. Vidal. Effect of silica particle size on polymer adsorp-
tion. Morphological, energetic and conformational relationships. Macromol.
Symp., 221, 271–280, 2005.
30. R. Iuliucci, C. Taylor, W. Kirk Hollis. 1H/29Si cross-polarization NMR experi-
ments of silica-reinforced polydimethylsiloxane elastomers: probing the
­polymer-filler interface. Magn. Reson. Chem., 44 (3), 375–384, 2006.
31. J.P. Cohen-Addad, P. Huchot, P. Jost, A. Pouchelon. Hydroxyl or methyl termi-
nated poly(dimethylsiloxane) chains: kinetics of adsorption on silica in mechan-
ical mixtures. Polymer, 30, 143–146, 1989.
32. J.P. Cohen-Addad, R. Ebengou. Silica-siloxane mixtures. Investigations into
adsorption properties of end-methylated and end-hydroxylated chains. Polymer,
33, 379–383, 1992.
33. J.P. Cohen-Addad, L. Dujourdy. Silica concentration dependence of the kinetics
of polydimethylsiloxane adsorption on aggregates. Polymer Bull., 41, 253–260,
1998.
34. L. Dujourdy. Mélanges de polydiméthylsiloxane et particules de silice.
Adsorption compétitive et cinétique. PhD Thesis, University of Grenoble 1,
Saint-Martin-d’Hères, France, 1996.
35. J.V. DeGroot, Jr., C.W. Macosko. Aging phenomena in silica-filled polydimethyl-
siloxane. J. Colloid Interface Sci., 217 (1), 86–93, 1999.
304 Filled Polymers

36. J.P. Cohen-Addad. Silica-siloxane mixtures. structure of the adsorbed layer: chain
length dependence. Polymer, 30, 1820–1823, 1989; Sol or gel-like behaviour of
ideal silica-siloxane mixtures: percolation approach. ibid., 33, 2762–2767, 1992.
37. R.S. Ziegelbaur, J.M. Caruthers. Rheological properties of poly(dimethylsiloxane)
filled with fumed silica. I. Hysteresis behaviour. J. Non-Newtonian Fl. Mech., 17,
45–68, 1985.
38. L.E. Kosinski, J.M. Caruthers. Rheological properties of PDMS filled with
fumed silica. II. Stress relaxation and stress growth. J. Non-Newtonian Fl. Mech.,
17, 69–89, 1985.
39. L.E. Kosinski, J.M. Caruthers. The effect of particle concentration on the rheol-
ogy of polydimethylsiloxane filled with fumed silica. J. Appl. Polym. Sci., 32,
3393–3406, 1986.
40. M.I. Aranguren, E. Mora, J.V. Degroot, C.W. Macosko. Effect of reinforcing fill-
ers on the rheology of polymer melts. J. Rheol., 36, 1165–1182, 1992.
41. M.I. Aranguren, E. Mora, C.W. Macosko. Compounding fumed silicas into poly-
dimethylsiloxane: Bound rubber and final aggregate size. J. Coll. Interf. Sci., 195
(2), 329–337, 1997.
42. S. Selimovic, S.M. Maynard, Y.Hub. Aging effects of precipitated silica in
poly(dimethylsiloxane). J. Rheol., 51, 325–340, 2007.
43. S.E. Shim, A.I. Isayev. Rheology and structure of precipitated silica and
poly(dimethyl siloxane) system. Rheol. Acta, 43 (2), 127–136, 2004.
44. P. Vondráček, M. Schätz. Bound rubber and crepe hardening in sulicone rubber.
J. Appl. Polym. Sci., 21, 3211–3222, 1977.
45. L. Léger, H. Hervet, M. Deruelle. Adsorption of polydimethylsiloxane chains on
plane silica surfaces. In Adsorption on Silica Surfaces, ed. E. Papirer. Dekker, New
York, NY, 597–619, 2000.
46. K. Polmanteer. Silicone rubber, its development and technological progress.
Rubb. Chem. Technol., 61, 470–502, 1988.
47. F. Clément. Etude des mécanismes de renforcement dans les réseaux de polydi-
methylsiloxane chargés silice. PhD Thesis, University Paris, France, 1999.
48. W.H. Waddell, L.R. Evans. Use of nonblack fillers in tire compounds. Rubb.
Chem. Technol., 69, 377–423, 1996.
49. Q. Liua, Y. Zhangb, H. Xuc. Properties of vulcanized rubber nanocomposites filled
with nanokaolin and precipitated silica. Appl. Clay Sci., 42 (1–2), 232–237, 2008.
50. R.J. Eldred. Effect of oriented platy filler on the fracture mechanism of elasto-
mers. Rubb. Chem. Technol., 61, 620–629, 1988.
51. H. Mouri and K. Akutagawa. Improve wet traction through the use of mineral
fillers. Rubb. Chem. Technol., 72, 960–968, 1999.
52. O. Noel, G. Meli. Effect of talc on rolling resistance in tread compounds. AFICEP
Technical Day, Paris, France, Dec. 3, 2008. Paper #1.
53. S. Yariv. Wettability of clay minerals. In Modern Approaches to Wettability: Theory
and Applications, M.E. Schrader and G. Loeb, Eds. Plenum Press, New York, NY,
279–326, 1992.
54. T.S. Stephens, H.H. Winter, M. Gottlieb. The steady shear viscosity of filled poly-
meric liquids described by a linear superposition of two relaxation mechanisms.
Rheol. Acta, 27, 263–272, 1988.
55. A.L. Graham, R.B. Bird. Particle clusters in concentrated suspensions.
1  Experimental observations of particle clusters. Ind. Eng. Chem. Fundam., 23,
406–410, 1984.
Polymers and White Fillers 305

56. A.L. Graham, R.D. Steele. Particle clusters in concentrated suspensions. 2.


Information theory and particle clusters. Ind. Eng. Chem. Fundam., 23, 411–420,
1984.
57. V. Vand. Viscosity of solutions and suspensions. I. Theory. J. Phys. Chem., 52 (2),
277–299, 1948.
58. D.J. Jeffrey, A. Acrivos. The rheological properties of suspensions of rigid parti-
cles. AIChE J., 22 (3), 417–432, 1976.
59. A.B. Metzner. Rheology of suspensions in polymeric liquids. J. Rheol., 29, 739–
775, 1985.
60. H. Eilers. Die Viskosität von Emulsionen hochviskoser Stoffe as Function der
Konzentration. Kolloid Z., 97, 913–321, 1941.
61. M. Mooney. The viscosity of a concentrated suspension of spherical particles. J.
Colloid Sci., 6, 162–170, 1951.
62. I.M. Krieger, T.J. Dougherty. A mechanism for non-Newtonian flow in suspen-
sions of rigid spheres. Trans. Soc. Rheol., 3, 137–152, 1959.
63. R. Ball, P. Richmond. Dynamics of colloidal dispersions. J. Phys. Chem. Liquids, 9,
99–116, 1980.
64. T. Kitano, T. Kataoka, T. Shirota. An empirical equation of the relative viscosity of
polymer melts filled with various inorganic fillers. Rheol. Acta, 20, 207–209, 1981.
65. A.L. Graham, R.D. Steele, R.B. Bird. Particle clusters in concentrated suspen-
sions. 3. Prediction of suspension viscosity. Ind. Eng. Chem. Fundam., 23, 420–425,
1984.
66. A.L. Graham. On the viscosity of suspensions of solid spheres. Appl. Sci. Res., 37,
275–286, 1981.
67. G.K. Batchelor. The effect of Brownian motion on the bulk stress in a suspension
of spherical particles. J. Fluid Mech., 83, 97–117, 1977.
68. J.S. Chong, E.B. Christiansen, A.D. Baer. Rheology of concentrated suspensions.
J. Appl. Polym. Sci., 15, 2007–2021, 1971.
69. A.J. Poslinki, M.E. Ryan, R.K. Gupta, S.G. Seshadri, F.J. Frechette. Rheological
behavior of filled polymeric systems. II. The effect of a bimodal size distribution
of particulates. J. Rheol., 32, 751–771, 1988.
70. A.J. Poslinki, M.E. Ryan, R.K. Gupta, S.G. Seshadri, F.J. Frechette. Rheological
behavior of filled polymeric systems. II. Yield stress and shear thinning effects.
J. Rheol., 32, 703–735, 1988.
71. M. Gahleitner, K. Bernreitner, W. Neißl. Correlations between the rheological
and mechanical properties of mineral filled polypropylene compounds. J. Appl.
Polym. Sci., 53, 283–289, 1994.
72. J.A. Radosta. Impact and flexural modulus behaviour of calcium carbonate and
talc filled polyolefins. Proceedings 33rd SPE-ANTEC, 525–530, 1975.
73. J. Jancar, A.T. DiBenedetto, A. Dianselmo. Effect of adhesion on the fracture
thoughness of calcium carbonate-filled polypropylene. Polym. Eng. Sci., 33, 559–
563, 1993.
74. B. Haworth, C.L. Raymond. Processing and fracture characteristics of polypro-
pylene filled with surface modified calcium carbonate. Proceedings Eurofillers 97,
Manchester, UK, Sept. 8–11, 251–254, 1997.
75. Y.W. Leong, M.B. Abu Bakar, Z.A. Mohd. Ishak, A. Ariffin, B. Pukanszky.
Comparison of the mechanical properties and interfacial interactions between
talc, kaolin, and calcium carbonate filled polypropylene composites. J. Appl.
Polym. Sci, 91, 3315–3326, 2004.
306 Filled Polymers

76. S. Wu. A generalized criterion for rubber toughening: the critical matrix liga-
ment thickness. J. Appl. Polym. Sci., 26, 1855–1863, 1985.
77. O.K. Muratoglu, A.S. Argon, R.E. Cohen, M. Weinberg. Toughening mechanism
of rubber-modified polyamides. Polymer, 36, 921–930, 1995.
78. P.M. McGenity, J.J. Hooper, C.D. Paynter, A.M. Riley, C. Nutbeem, N.J. Elton,
J.M. Adams. Nucleation and crystallization of polypropylene by mineral fillers:
relationship to impact strength. Polymer, 33, 5215–5224, 1992.
79. C.-M. Chan, J. Wu, Y.-K. Cheung, J. Li. Polypropylene/calcium carbonate nano-
composites. Polymer, 43, 2981–2992, 2002.
80. P. Supaphol , W. Harnsiri. Rheological and isothermal crystallization character-
istics of neat and calcium carbonate-filled syndiotactic polypropylene. J. Appl.
Polym. Sci., 100, 4515–4525, 2006.
81. C.D. Han, T. Van Den Weghe, P. Shete, J.R. Haw. Effects of coupling agents on
the rheological properties, processability, and mechanical properties of filled
polypropylene. Polym. Eng. Sci., 21, 196–204, 2004.
82. Y.W. Leong, M.B. Abu Bakar, Z.A. Mohd Ishak , A. Ariffin. Effects of filler treat-
ments on the mechanical, flow, thermal, and morphological properties of talc
and calcium carbonate filled polypropylene hybrid composites. J. Appl. Polym.
Sci., 98, 413–426, 2005.
83. Y. Wang , W.-C. Lee. Interfacial interactions in calcium carbonate-polypropylene
composites. 1: surface characterization and treatment of calcium carbonate: a
comparative study. Polym. Compos., 24, 119–131, 2004.
84. Y. Wang , W.-C. Lee. Interfacial interactions in calcium carbonate-polypropylene
composites. 2: effect of compounding on the dispersion and the impact proper-
ties of surface-modified composites. Polym. Compos., 25, 451–460, 2004.
85. D.M. Ansari, G.J. Price. Correlation of the material properties of calcium carbon-
ate filled polypropylene with the filler surface energies. J. Appl. Polym. Sci., 88,
1951–1955, 2003.
86. G.J. Price, D.M Ansari. Surface modification of calcium carbonates studied by
inverse gas chromatography and the effect on mechanical properties of filled
polypropylene. Polym. Intern., 53, 430–438, 2004.
87. C.D. Han, C. Sandford, H.J. Yoo. Effects of titanate coupling agents on the rheo-
logical and mechanical properties of filled polyolefins. Polym. Eng. Sci., 18, 849–
854, 2004.
88. K.-J. Kim, J.L. White, S.E. Shim, S. Choe. Effects of stearic acid coated talc,
CaCO3, and mixed talc/CaCO3 particles on the rheological properties of poly-
propylene compounds. J. Appl. Polym. Sci., 93, 2105–2113, 2004.
89. Y. Wang, J.-J. Wang. Shear yield behavior of calcium carbonate-filled polypro-
pylene. Polym. Eng. Sci., 39, 190–198, 2004.
90. C. Richard, K. Hing, H.P. Schreiber. Interaction balances and properties of filled
polymers. Polym. Compos., 6, 201–208, 2004.
91. T.L. Smith. Volume changes and dewetting in glass bead–polyvinyl chloride
elastomeric composites inder large deformations. Trans. Soc. Rheol., 3, 113–136,
1959.
92. L.E. Nielsen. Simple theory of stress-strain properties of filled polymers. J. Appl.
Polym. Sci., 10, 97–103, 1966.
93. J. Bohse, S. Grellmann, S. Seidler. Micromechanical interpretation of fracture
toughness of particulate-filler thermoplastics. J. Mat. Sci., 26, 6715–6721, 1991.
Polymers and White Fillers 307

94. H.E. Wiebking. The performance of ultrafine talc in rigid PVC. J. Vinyl Additive
Technol., 2 (3), 187–189, 1996.
95. D.W. Cornwell. Plastic performance: benefits of PCC as a PVC additive. Ind.
Min., July, 35–37, 2001.
96. G. Guerrica-Echevarria, J.I. Eguiazabal, J. Nazabal. Influence of molding con-
ditions and talc content on the properties of polypropylene composites. Eur.
Polym. J., 34, 1213–1219, 1998.
97. B. Pukanszky, K. Belina; A. Rockenbauer, F. Maurer. Effect of nucleation, filler
anisotropy and orientation on the properties of PP composites. Composites, 25,
205–214, 1994.
98. W. Qiu, K. Mai, H. Zeng. Effect of silane-grafted polypropylene on the mechani-
cal properties and crystallization behavior of talc/polypropylene composites. J.
Appl. Polym. Sci., 77, 2974–2977, 2000.
99. A.L.N. da Silva, M.C.G. Rocha, M.A.R. Moraes, C.A.R. Valente, F.M.B. Coutinho.
Mechanical and rheological properties of composites based on polyolefin and
mineral additives. Polym. Test., 21, 57–60, 2002.
100. C.D. Han, C. Sandford, H.J. Yoo. Effects of titanate coupling agents on the rheo-
logical and mechanical properties of filled polyolefins. Polym. Eng. Sci., 18, 849–
854, 2004.
101. Y.W. Leong, M.B. Abu Bakar, Z.A. Mohd. Ishak, A. Ariffin. Effects of filler treat-
ments on the mechanical, flow, thermal, and morphological properties of talc
and calcium carbonate filled polypropylene hybrid composites. J. Appl. Polym.
Sci., 98, 413–426, 2005.
102. S. Fellahi, N. Chikhi, M. Bakar. Modification of epoxy resin with kaolin as a
toughening agent. J. Appl. Polym. Sci., 82, 861–878, 2001.
103. G. Qiu, F. Raue, G.W. Ehrenstein. Mechanical properties and morphologies of
PP/mPE/filler composites. J. Appl. Polym. Sci., 83, 3029–3035, 2001.
104. Q.H. Zeng, A.B. Yu, G.Q. (Max) Lu, D.R. Paul. Clay-based polymer nanocom-
posites: Research and commercial development. J. Nanosci. Nanotechnol., 5,
1574–1592, 2005.
105. S.-Y. Fu, X.-Q. Feng, B. Lauke, Y.-W. Mai. Effects of particle size, particle/matrix
interface adhesion and particle loading on mechanical properties of particulate–
polymer composites. Composites: Part B, 39, 933–961, 2008.
308 Filled Polymers

Appendix 6

A6.1  Adsorption Kinetics of Silica on Silicone Polymers


Model of polymer adsorption kinetics

Q(t) = Q0 + Qinf ( 1 − exp ( − λ ⋅ t )) Qinf: quantity of adsorbed polymer for an


infinite contact time
Q 0: initial quantity of adsorbed polymer
(i.e., immediately after mixing)
λ: kinetic constant for the adsorption
­process (λ = tad–0.5 with tad a characteristic
time)

A6.1.1 Effect of Polymer Molecular Weight


PDMS Mn (g/mole): [Data: L. Dujourdy, PhD Thesis, University
J. Fourier, Grenoble, France, 1996]

43,000 73, 000 300,000

 0 0.264   0 0.405   0 0.775 


 175
 0.48   28
 0.588 
  70
 1.197 
 525 0.673   42 0.588   238 1.69 
     
1015 0.704   245 1.092  595 1.901 
Data1 :=  Data3 : = 
1715 0.722   245 1.099  1085 2.077 
        
 2293 0.792   413 1.197  1918 2.113 
 2713 0.785  Data2: =  413 1.25   2415 2.116 
  
 3308 0.792   840 1.3   3255 2.134 
 
time Q(t)  1190 1.498 
h g/g silica
 1838 1.567 
 
 2030 1.563 
 2695 1.482 
Silica: 150 m2/g  
 2842 1.514 
Weight fraction: 0.20
Polymers and White Fillers 309

Model equation and


partial derivatives
Collecting experimental data
 C0 + C1 ⋅ ( 1 − exp ( − C2 ⋅ x ))
j : = 0 .. 2 : datasets  
 1 
F(x, C): = 
x0 : = Data1<0> y 0 : = Data1<1> 1 − exp ( − C2 ⋅ x ) 
 
 C1 ⋅ x ⋅ exp ( − C2 ⋅ x ) 
x1 : = Data2<0> y1 : = Data2<1>

x2 : = Data3<0> y 2 : = Data3<1> nj : = length ( x j )

 (yj ) 
 0
 : initial guessed parameters
C j : = ( y j ) n − 1 
 j

 0.1 

Calling the nonlinear


Resj: = GenFit(xj,yj,Cj,F)
   fitting algorithm:  

Fa j : = ( Res j )0 + ( Res j )1 ⋅ 1 − exp  − ( Res j )2 ⋅ x j     : fit equation         Characteristic time

R2j: = corr(Faj,yji) : correlation coefficient 1


t ad j : =
( Res j ) 
2

2

Fitting data on PDMS (43,000 g/mol)/silica compound


Adsorbed polymer, g/g of filler

 0.257 
2
Res 0 =  0.588 
 0.044 
tad = 514.819 1
0

R2 0 = 0.991 (Res 0 )0 + (Res 0 )1 = 0.845

  0
0 1000 2000 3000
Time, h
310 Filled Polymers

Fitting data on PDMS (73,000 g/mol)/silica compound

Adsorbed polymer, g/g of filler


 0.32 
Res1 =  1.321 
2

 0.053 

t ad1 = 350.675 1

R2 1 = 0.987 (Res1 )0 + (Res1 )1 = 1.641

0
0 1000 2000 3000
Time, h

Fitting data on PDMS (300,000 g/mol)/silica compound


Adsorbed polymer, g/g of filler

 0.733 
Res 2 =  1.483 
2

 0.061

t ad2 = 273.138 1
R2 2 = 0.992 (Res 2 )0 + (Res 2 )1 = 2.217

0
0 1000 2000 3000
Time, h

A6.1.2 Effect of Silica Weight Fraction


Silica weight fraction: [Data: L. Dujourdy, PhD Thesis, University J. Fourier,
Grenoble, France, 1996]
0.049 0.103 0.204
 0 0.206   0 0.265 
 38  0 0.279 
0.312   163
 184 0.456 

 0.55  
 61 0.497   510 0.676 
   653 0.691   
163 0.676    1000 0.691
Data4: =  Data5: =  980 0.779  Data6: = 
 469 1.241  1706 0.721
   1306 0.882   
 1388 1.5     2245 0.788 
 2000   2286 0.985   2673
1.588  3633 0.794 
   1   
 3327 1.544   3265 0.794 

PDMS, Mn = 43,000 g/mole


Silica: 150 m2/g
Polymers and White Fillers 311

j : = 3..5 : datasets

x3 : = Data4<0> y 3 : = Data4<1>
: collecting experimental data
x4 : = Data5<0> y 4 : = Data5<1>
n j : = length( x j )
x5 : = Data6<0> y 5 : = Data6<1>

 ( y j )0  : initial guessed parameters


 
C j : = ( y j )n j −1   
 
 0.1 

Calling the nonlinear    Res : = GenFit(x ,y ,C ,F)


j j j j
fitting algorithm:

Fa j : = ( Res j )0 + ( Res j )1 ⋅ 1 − exp  − ( Res j )2 ⋅ x j     : fit equation      Characteristic time

R2j: = corr (Faj,yj) : correlation coefficient  t : = 1


( Res j ) 
ad j 2

2

Fitting data on 0.049 silica weight fraction compound

2
Adsorbed polymer, g/g of filler

 0.103 
Res 3 = 1.702 
 0.041
1
tad 3 = 588.735

R2 3 = 0.98 (Res 3 )0 + (Res 3 )1 = 1.805

0
0 2000 4000
Time, h

312 Filled Polymers

Fitting data on 0.103 silica weight fraction compound

Adsorbed polymer, g/g of filler


 0.266 
Res 4 = 1.042 
 0.022  1

tad4 = 2.02 ⋅ 10 3

R2 4 = 0.993 (Res 4 )0 + (Res 4 )1 = 1.308


0
0 2000 4000
Time, h

Fitting data on 0.204 silica weight fraction compound

2
Adsorbed polymer, g/g of filler

 0.255 
Res 5 =  0.614 
 0.039  1

tad5 = 648.22

R2 5 = 0.985 (Res 5 )0 + (Res 5 )1 = 0.869


0
0 2000 4000
Time, h

A6.2 Modeling the Shear Viscosity Function of Filled


Polymer Systems
By combining two Carreau–Yasuda equations, an eight parameter model is
obtained that would meet all the likely typical features of the shear viscosity
behavior of filled polymer systems.

n 1− 1 n 2 -1

η(γ °) = η 01⋅ 1 + ( λ 1⋅ γ °) 1 
a a1
+ η 02⋅ 1 + ( λ 2⋅ γ °) 2  a 2
a  he (very) low shear
T
   
plateau is accounted for
nonhydrodynamic interparticless by parameter η01 and the
effectsof filler hydrodynamic    high shear thinning region
is depending on the flow
particles effectts index n2.
Polymers and White Fillers 313

Nonhydrodynamic effects of filler particles (e.g., filler networking)


­ ominate the low shear behavior in such a manner that an apparent yield-
d
ing region overrides the pseudo-Newtonian plateau of the polymer matrix
­(corresponding to parameter η02)
There is an intermediate plateau when:

n 1− 1
η 02 = η c = η 01⋅ [1 + (λ 1⋅ γ °cc )a 1 ] a1

which corresponds to a critical shear rate γc:

 1
  a1 
 a 1 
1  η 02   n 1−1  
γ °c = ⋅ −1
λ 1  η 01  
 

The critical shear rate γc corresponds to a critical characteristic time λc


A few mathematical aspects of the model:
γ°: = 0.0001⋅sec−1, 0.001⋅sec−1 .. 200⋅sec−1 : shear rate range for calculations
Model parameters:

1
η01 : = 8 ⋅ 10 4 ⋅ Pa ⋅ sec λ 1 : = 500 ⋅ sec a1 : = 1.9 n1 : = 0.4 γ °c l : =
λ1

1
η02 : = 3 ⋅ 103 ⋅ Pa ⋅ sec λ 2 : = 0.1 ⋅ sec a2 := 3 n2 := 0.33 γ °c 2 : =
λ2

n 1− 1 n 2−1

η(γ ° ) : = η 01⋅ 1 + ( λ 1⋅ γ °
 ) a1


a1
+ η 02⋅ 1 + ( λ 2⋅ γ °
 ) a2


a2
: Model equation

γ ° c 1 = 2 ⋅ 10−3°sec −1 η( γ ° c 1 ) = 6.727 ⋅ 10 4 °Pa ⋅ sec

γ ° c 2 = 2 × 10°sec −1 η( γ ° c 2 ) = 3.052 ⋅ 103 °Pa ⋅ sec

 1
  a1 
 a 1 
1  η 02   n 1−1  
γ °c : = ⋅ −1
Critical shear rate corresponding λ 1  η 01  
to the intermediate plateau:  

γ °c = 0.476 ° sec −1 η c : = η ( γ °c ) η c = 6.103 °Pa ⋅ sec   :viscosity at the ­critical


shear rate
314 Filled Polymers

η01
1.105 : Shear viscosity function
γ°c1 γ°c2 with respect to ­parameters
Shear viscosity, Pa.s

η01 and η02.


1.104
η02

1.103

100
1.10–4 1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s

 hear viscosity function


S 1.105
with respect to critical shear
Shear viscosity, Pa.s

γ°c
rate and viscosity: ηc
1.104

1.103

100 –4
1.10 1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s

n 1− 1 n 2−1

η(γ °, n1 ) : = η 01⋅ 1 + ( λ 1⋅ γ °) 1  + η 02⋅ 1 + ( λ 2⋅ γ °) 2 


a a
n1 : = 0.1,0.2 .. 0.5 a1 a2
   

1.105

    : Effect of parameter n1


Shear viscosity, Pa.s

1.104

1.103

100
1.10–4 1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
n1 = 0.2 n1 = 0.5

n 1− 1 n 2 −1
a 1: = 0.5, 1.6 .. 3.5 n 1: = 0.3 η(γ º , a 1) : = η 01⋅ 1 + ( λ 1⋅ γ º ) 1  + η 02⋅ [ 1 + (λ 2⋅ γ ° )a 2 ]
a a1 a2
 
Polymers and White Fillers 315

1.105
    : Effect of parameter a1
Shear viscosity, Pa.s

1.104

1.103

100
1.10–4 1.10–3 0.01 0.1 1 10
Shear rate, 1/s
a1 = 0.8 a1 = 3.2

A6.3 Models for the Rheology of Suspensions of Rigid


Particles, Involving the Maximum Packing Fraction Φm
The general approach consists in considering that the viscosity of the sus-
pension is equal to the viscosity of the suspending medium times a suitable
functional whose mathematical form is derived from theoretical consider-
ations. No specifc interactions are considered between the particles and the
suspending medium; only nonhydrodynamic crowding effects are consid-
ered in such a manner that the viscosity of the systems goes to infinity as
the particle volume fraction approaches the maximum packing fraction of
the particles.
Besides the theoretical reasonning in deriving the model equation, there
are two criteria for validity:

1. As the volume fraction decreases, the viscosity must become asymp-


totic to the straightline of slope 2.5 corresponding to the Einstein’s
equation
2. As the volume fraction increases and approaches the maximum
packing fraction, the viscosity must go to infinity

All the models considered below meet those two validity criteria.

Φ: = 0,0.01 .. 0.5 : volume fraction range for calculations


π
Φ m: = : maximum packing fraction of particles (here,
3⋅ 2 spheres of equal diameter in the closest hexagonal
packing)
316 Filled Polymers

ηE(Φ): = (1 + 2.5⋅Φ) : Einstein (1906–1911); low volume fraction limit

2
 5 
1+ ⋅Φ 
ηEil( Φ ) :=  2  : Eilers (1941) Guth, Gold and Simha (1948)
 1− Φ  ηGGS(Φ): = (1 + 2.5⋅Φ + 14.1⋅Φ2)
 Φ m 

 
5 Φ 
η M(Φ): = exp  ⋅  : Mooney (1951)
2 1− Φ 
 Φm 

−5
⋅Φ m
 Φ  2
ηKD( Φ ) : =  1 − : Krieger–Dougherty (1959)
 Φ m 

−2
 Φ 
ηK ( Φ ) : =  1 − : Kitano et al. (1981)
 Φ m 

  : Graham (1981)
 1 
 3 3 2 
 1− Φ 
1 −
Φ  
 Φ  Φ   Φ  
5 9
ηG( Φ ) : = 1 + ⋅ Φ + ⋅  2 ⋅ 3 Φ
m
⋅  1 +  ⋅ 1 + 2 ⋅ 3 m  
 2  Φ  
2 4 
 Φm  Φm  
 
 
 
 
 

−2.5
   1− Φ   Φ m− Φ   
2

ηGSB (Φ) : = 1 − Φ ⋅ 1 +  m
 ⋅ 1 −  Φ   : Graham et al. (1984)
   Φ m  m    
Polymers and White Fillers 317

Comparing models

Models predicting Models predicting


a sharp variation vs. Φ a smooth variation vs. Φ

50 20

40
15
Relative viscosity

Relative viscosity
30
10
20

5
10

0 0
0 0.2 0.4 0 0.2 0.4
Volume fraction Volume fraction
Einstein Eilers Einstein Krieger & Dougherty
Mooney Kitano Graham Graham, Steele & Bird
Guth, Gold & Simha Guth, Gold & Simha
  

Effect of maximum packing fraction


Φm: = 0.5 .. 0.74 [lower bound: cubic packing = 0.524]
[higher bound: hexagonal packing = 0.74]

 
5 Φ 
Mooney (1951): η M( Φ, Φ m ): = exp  ⋅ 
2 1− Φ 
 Φm 
100
Relative viscosity

50

0
0 0.2 0.4 0.6 The larger Φm, the larger the
Volume fraction ­ olume fraction range before
v
Max Fract. = 0.524 Max Fract. = 0.6 ­relative ­viscosity goes to infinity.
Max Fract. = 0.7

318 Filled Polymers

−5
⋅Φ m
 Φ  2
η KD( Φ , Φ m ): =  1 −
Krieger–Dougherty (1959):  Φ m 

40
Relative viscosity

20

0
0 0.2 0.4
Volume fraction  he larger Φm, the smoother
T
Max Fract. = 0.524 Max Fract. = 0.6 the variation in relative
Max Fract. = 0.7 viscosity.

 
 
 
 1 
 2 
5 9  3
Φ  3 Φ
 
Graham (1981): ηG ( Φ , Φ m ) : = 1 + ⋅ Φ + ⋅ 1−  1−  
2 4  Φm  Φ Φm
2⋅ 3 ⋅  1 +  ⋅ 1 + 2 ⋅ 3  
 Φ  2  Φ  
 Φm  Φm  
 
 
 

40
Relative viscosity

20

0
0 0.2 0.4
Volume fraction The larger Φm, the smoother
Max Fract. = 0.524 Max Fract. = 0.6 the  variation in relative

Max Fract. = 0.7 viscosity.
Polymers and White Fillers 319

−2.5
   1− Φ   Φ − Φ  
2

Graham et al.: η GSB(Φ, Φ m): = 1 − Φ ⋅ 1 +  m


 ⋅ 1−  m  
   Φ m   Φ m  


40
Relative viscosity

20

0  he larger Φm, the larger


T
0 0.2 0.4
Volume fraction the volume fraction range
Max Fract. = 0.524 Max Fract. = 0.6 before relative viscosity
Max Fract. = 0.7 goes to infinity.

A6.4 A
 ssessing the Capabilities of the Model for the
Shear Viscosity Function of Filled Polymers
γ: = 0.001 sec.−1, 0.01 sec−1...1000 sec−1 : shear rate range for calculations
Φ: = 0, 0.01 .. 0.5 : volume fraction range for calculations
Φm: = 0.64 : maximum packing fraction for a random arrangement
of spheres (of equal diameter) 1
= 100 °sec −1
η0: = 3 ⋅ 103 ⋅ Pa ⋅ sec : zero-shear viscosity of polymer matrix λ 0
λ0: = 0.01 ⋅ sec : characteristic time of polymer matrix
n: = 0.3 : flow index (in high shear region)
a: = 2 : Yasuda exponent (curvature in transition region for
the polymer matrix)
σc: = 200 ⋅ Pa : yield stress of filled compound

n−1

  −2  −2
 
a a
σ Φ      Φ    : model equation
η( Φ ,γ  ,λ 0 , a,σ c ) : = c +  η 0⋅  1 −  ⋅ 1 + λ
 0 ⋅ 1 − ⋅ γ 
 
γ   Φ m      Φ m   

320 Filled Polymers

A6.4.1 Effect of Filler Fraction


(λ0 cst = 0.01 s, a = 2, σc = 200 Pa)
1.106

Shear viscosity, Pa.s


1.105

1.104

1.103

100
1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
Vol.Fract.=0.1 Vol.Fract.=0.3
Vol.Fract.=0.5

A6.4.2 Effect of Characteristic Time λ0


(Φ cst = 0.2, a = 2, σc = 200 Pa)
1.106

1.105
Shear viscosity, Pa.s

1.104

1.103

100

10
1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
Lambda = 0.01 Lambda = 0.05
Lambda = 0.6
Polymers and White Fillers 321

A6.4.3 Effect of Yasuda Exponent a


(Φ cst = 0.3, λ0 cst = 0.01 s, σc = 200 Pa)
1.106

Shear viscosity, Pa.s


1.105

1.104

1.103

100
1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
a = 0.5 a = 1.0 a = 2.0

A6.4.4 Effect of Yield Stress σc


(Φ cst = 0.3, λ0 cst = 0.01 s, a = 2)
1.106
Shear viscosity, Pa.s

1.105

1.104

1.103

100
1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
Yield stress = 100 Pa
Yield stress = 250 Pa
Yield stress = 500 Pa
322 Filled Polymers

A6.4.5  Fitting Experimental Data for Filled Polymer Systems


Fitting experimental data with the above model is not straightforward and
requires both an extanded set of data and the appropriate fitting strategy,
as demonstrated below with shear viscosity data on a series of TiO2 filled
polystyrene at 180°C, as published by Minagawa and White [J. Appl. Polym.
Sci, 20, 501–523, 1976].
In order to apply the model, TiO2 particles are considered as spheres of
equal diameter in such a manner that the maximum packing fraction Φm can
be considered equal to 0.64.

virgin Polystyrene (PS) PS + 4.8% TiO 2 PS + 14.3% TiO 2

Φ TiO = 0.012 Φ TiO = 0.034


2 2

Shear Shear
Rate Shear Shear Shear Shear
Visco Rate Visco Rate Visco
s−1 kPa.s s-1 kPa.s s−1 kPa.s

 0.0027 17.191   0.0012 22.701   0.0012 53.022 


 0.00072 15.781   0.0027 21.750   0.0027 41.314 
     
 0.0258 15.781   0.0072 19.544   0.0072 32.191 
 0.0798 13.684   0.0258 18.860   0.0258 27.914 
     
 0.2470 12.7743 
 0.0798 18.200   0.0798 24.205 
 0.4049 11.450   0.2470 15.781   0.2470 19.544 
 0.7646 8.922   0.4049 13.684   0.4049 16.354 
     
Data1 : =  1.6629 6.028   0.7646 9.581   0.8806 1.450 
11
Data2 : = Data3: =
 3.6164 4.374   1.6629 6.708   1.6629 8.308 
     
 8.4407 2.656 
 3.6164 5.044   3.6164 5.417 
 18.3571 1.731   8.4407 3.289   8.4407 3.532 
 39.9234 1.089
  18.3571   18.3571 
  1.997 2.303
   
 93.1805 0.638   39.9234 1.212   39.9234 1.501 
 233.3986 0.361   93.1805 0.736   93.1805 0.849 
     
 440.7292 0.219   2333.3986 0.416   233.3986 0.4480 
 440.7292 0.253   440.7292 0.291 
Polymers and White Fillers 323

PS + 24.0% TiO 2 PS + 38.6% TiO 2

Φ TiO2 = 0.056 Φ TiO2 = 0.087

Shear Shear
Shear Shear  0.0027 393.422 
 0.0072 178.145  Rate Visco
Rate Visco
  s−1 kPa.s
s -1 kPa.s
 0.0146 100.715 
   0.8806 25.083 
 0.0258 70.518   2.0553 
14.695
 0.0422 56.940   
   3.6164 9.581 
 0.0798 422.813   
 0.1308 34.570   9.0585 5.613 
   18.3571 3.793 
 0.2470 27.914  Data5 : =  
   39.9234 2.473 
0.4049 23.357 
Data4 : =   93.1805 1.398 
 0..8806 15.781   
   195.1805 1.398 
 1.7846 9.929   
 3.6164 6.708   195.6190 0.790 
   440.77292 0.480 
 9.0585 73 
4.07
 
 18.3571 2.752 
 39.9234 1.859 
 
 93.1805 1.051 
 195.66190 0.594 
 
 440.7292 0.348 

A6.4.6  Observations on Experimental Data

1.103
Shear viscosity, kPa.s

100

10
: Experimental shear
viscosity functions
1
0.1
1.10–3 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
Virgin PS 4.8% TiO2 14.3%
18.0% 38.6%
324 Filled Polymers

As can be seen, the shear viscosity function is well investigated over six
decades of shear rate for the virgin polystyrene and the filled compounds
up to 18% TiO2 content. Whilst the 38.6% compound is only documented
in a shorter shear rate range (i.e., 1–1000 s−1), the occurrence of a yield stress
behavior that depends on the filler level makes however no doubt.
In the high shear region, the shear thinning behavior is not much affected
by the presence of the filler, since the flow index n seems to be the same,
whatever is the filler content. Presumably, the characteristic time λ0 is also
not depending on the filler content.
It follows that, for the filled compounds, only σc η0 and a must be fit to
experimental data

A6.4.7 Extracting and Arranging Shear Viscosity Data

j: = 0 .. 4

γ 10 : = Data1< 0> γ 11 : = Data2 < 0> γ 12 : = Data3< 0> γ 13 : = Data 4< 0> γ 1 4 : = Data5< 0>

η0 : = Data1< 1> η1 : = Data2 < 1> η2 : = Data3< 1> η3 : = Data 4< 1> η4 : = Data5< 1>

n1: = length ( γ 1 0 )
 n2 : = length ( γ 1 1 )
 n3 : = length ( γ 1 2 )
 n4 : = length ( γ 1 3 )
 n5 : = length ( γ 1 4 )

n1 = 15 n2 = 16 n3 = 16 n4 = 18 n5 = 9

A6.4.8 Fitting the Virgin Polystyrene Data


with the Carreau–Yasuda Model
n− 1

η1( γ °,λ 0, a ) = η 0⋅ 1 + ( λ 0⋅ γ °) 
a a
: Carreau–Yasuda equation
 

Guess parameters for nonlinear fitting algorithm


  (GenFit function)
 (η ) + (η ) 
0 0 0 2
 
 2  n1: = length (η0) : number of data
   < = extracting guess parameters from experi-
 1  mental data, i.e.,
C: =  
( γ 1°0 )round  n1 , 0 -T  he average of two lowest shear rate viscosity
  2  
  data for η0
 0.3  - The reverse of the mid shear rate range for λ0
  - A common value for molten polymers for the
 2 
  flow index
- a = 2 (Carreau model)
Polymers and White Fillers 325

Model equation and partial


derivatives:

 C2 − 1

C0 ⋅ 1 + ( C1 ⋅ γ ° ) 3  C3
C
   
 
 C2 − 1 
 1 + ( C1 ⋅ γ ° )C3  C3 
   
 
 C0
 C2 − C3 − 1  
⋅ ( C2 − 1) ⋅ ( C1 ⋅ γ ° ) 3 ⋅ 1 + ( C1 ⋅ γ ° ) 3  C3 
C C  
 
C   
F( γ °, C):=  1
 
C2 − 1
 C0  
⋅ 1 + ( C1 ⋅ γ ° ) 3  C3 ⋅ ln 1 + ( C1 ⋅ γ ° ) 3 
C C
     
 C 3 
 
  C3 ⋅ ( C1 ⋅ γ ° )C3 ⋅ ln ( C1 ⋅ γ ° ) 
  … 
 C2 − 1
 C3 ⋅ 1 + ( C1 ⋅ γ ° )C3   
   
  C2 − 1    C3 ⋅ 
C0 ⋅  C  ⋅ 1 + ( C1 ⋅ γ ° ) 
C3

 3  1 + ( C1 ⋅ γ ° ) 3  ⋅ ln 1 + ( C1 ⋅ γ °
C
)   
C3

 + −    
  C3 ⋅ 1 + ( C1 ⋅ γ ° )C3   
     

Res1: = GenFit ( γ 1°0, η0, C, F ) : calling the nonlinear regression algorithm

p : = 1..24 zl : = 0.001 γ s° p := 10log( zl) + p⋅0.25   : discrete shear rate range


for calculations

Res12 − 1
 
η0fp : =  Res10 ⋅ 1 + ( Res11 ⋅ γs° p )
 
Res13 Res13
 : Fitting equation
   


 Res12 − 1
 
r02 : = corr   Res10 ⋅ 1 + ( Res11 ⋅ γ 1°0 ) 3  Res13 , η0 
Res1

    : Correlation
  
coefficient r2
Virgin PS
100
16.712 
η0
 1.183 
10 Res1 =  
 0.277 
  
η0fp  
1  0.77 

0.1
1.10–3 0.01 0.1 1 10 100 1.103 1.104
γ1°0, γs°p r02 = 0.998
  
326 Filled Polymers

Assembling results

 0  Φ
R1:= A0 ← 0 166.712 
  η0
for i ∈ 0..3
 1.183  λ0
A1 + i ← Res 1i  
R1 =  0.277  n
A5 ← 0  0.77  a
A6 ← r 02  
 0  σc
A  0.998 
  r2

A6.4.9  Fitting the Filled Polystyrene Shear Viscosity Data


Model equation:
n− 1
−2  −2
 
a a
σc   Φ      Φ 
η( γ °,λ 0, a,σ c ) = +  η 0⋅ 1 −  ⋅ 1 +  λ 0⋅ 1 − ⋅ γ ° 
γ °   Φ m      Φ m   

n− 1
  
a a
σc  Φm 2    Φm 2
or: η( γ °,λ 0, a, σ c ) = +  η 0⋅  ⋅ 1 +  λ 0⋅ ⋅ γ ° 
γ °  ( Φ m− Φ )2    ( Φ m− Φ )2  

When compared with the virgin PS, the shear viscosity data for the PS + 4.8%
TiO2 compound are essentially slightly shifted upwards with a possible
modification of the curve in the transition region (which depends in fact on
the value of the parameter a), and the yield stress behavior is barely visible
at low shear rate.
This suggests that the characteristic time λ0 and the flow index n obtained
for the virgin polymer can be kept for the filled compounds, so that only
the parameters are σc, η0 and a must be determined by nonlinear fitting.
Φ m: = 0.64 : maximum packing fraction for a random arrangement of
spheres (of equal diameter)

PS + 4.8%TiO2 ΦTiO2 = 0.012 Φ1: = 10.012 Φ: = Φ1

λ0: = Res11 n: = Res12 : model parameters considered as


constant (i.e., depending on the
polymer matrix only)
Polymers and White Fillers 327

 0.001 
D1 : =  Res10   < = guest parameters:  1.10−3 
 Res13  - A low value for σc D1= 16.712  Φ m2
-T he low shear viscosity A: =
of the virgin PS for η0
 0.77  ( Φ m − Φ )2
-F or a, the value obtained
for the virgin PS  A = 1.039

        Model equation and partial


               derivatives:

n− 1
 D0 
+ ( D1 ⋅ A ) ⋅ 1 + ( A ⋅ λ 0⋅ γ  ) 2  D2
D
   
 γ ° 
 1 
 
 γ° 
G(γ ° , D) : =   ( n − 1)  
 D   
 A ⋅ 1+ ( λ 0⋅ A ⋅ γ °) 2   D2  
 
 
 ( n − 1)  
ln 1 + ( λ 0⋅ A ⋅ γ ° )   
D2
 ( n − 1) ⋅ D1 ⋅ A ln ( λ 0⋅ A ⋅ γ ° ) ⋅ ( λ 0⋅ A ⋅ γ ° )D
2

−  
D  
 ⋅ 1 + ( λ 0⋅ A ⋅ γ ° ) 2   D2  ⋅  
 D 2
   1 + ( λ 0⋅ A ⋅ γ ° ) 2
D
D2 
  

Res2: = GenFit ( γ 1°1,η1, D1, G )   : calling the nonlinear regression algorithm


n− 1
 Res2 Res2 2 Res2 
η1fp : =  0
+ ( Res2 1 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ s° p )
  2
   : Fitting equation
 γ s° p   


 n− 1
 
Res2 0
 + ( Res2 1 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ 1°1 ) 2  Res22 , η1    : Correlation coefficient r2
Res2
r12 := corr 
  γ 1°1    

PS + 4.8% TiO2
100

η1 10
 3.285 ⋅ 10−3 
η1fp
1 Res2 =  19.469 
 0.772 
0.1 –3
1.10 0.01 0.1 1 10 100 1.103 1.104
γ1°1, γs°p
r12 = 0.997

328 Filled Polymers

Assembling results

R 2 : = A0 ← Φ
 0.012  Φ
A1 ← Res 2 1  19.469  η
for i ∈1.. 2   0
 1.183  λ 0
A1 + i ← Res 1i  
R 2 =  0.277  n
A4 ← Res 2 2  0.772  a
A5 ← Res 2 0  
 3.285 ⋅ 10−3  σ 2
A6 ← r 12  0.997  r 2
 
A

PS + 14.3% TiO2 ΦTiO  = 0.034 Φ2: = 0.034 Φ: = Φ2


2

λ0: = Res11 n: = Res12

 <= guest parameters:
 Res2 0  -F  or σc, the value obtained for the lower filled compound
D2 : =  Res2 1  above
 Res2 2  - For η0, the value obtained for the lower filled compound
above
-F  or a, the value obtained for the lower filled compound
above

 3.285 ⋅ 10−3 
Φ m2
D2 =  19.469  A: = A = 1.115
 0.772  ( Φ m − Φ )2

call G(γ°, D) : model equation and partial derivatives

Res3: = GenFit ( γ 1°2 , η2 , D2 ,G ) : calling the nonlinear regression algorithm


n− 1
 Res3 Res32 Res3 
η2fp : =  0
+ ( Res31 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ s° p )  2  Fitting equation
 γ s° p   
Polymers and White Fillers 329


 n− 1
 
Res30
 + ( Res31 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ 1°2 ) 2  Res32 , η2    Correlation coefficient r2
Res3
r22 : = corr 
  γ 1°2    

PS + 14.3% TiO2
100
 0.026 
Res3 =  31.734 
η2
10
η2fp  0.526 
1

0.1
1.10–3 0.01 0.1 1 10 100 1.103 1.104
γ1°2, γs°p
r22 = 0.999

Assembling results

R 3 : = A0 ← Φ
 0.034  Φ
A1 ← Res 31  31.734  η
for i ∈1.. 2   0
 1.183  λ 0
A1 + i ← Res 1i  
R 3 =  0.277  n
A4 ← Res 32  0. 526  a
A5 ← Res 30  
 0.026  σ c
A6 ← r 22  0.999  r 2
 
A

PS + 24.0% TiO2 ΦTiO  = 0.056 Φ3: = 0.056 Φ: = Φ3


2

λ0: = Res11 n: = Res12

 <= guest parameters:
- For σc, the value obtained for the lower filled compound
 Res30  above
D3 : =  Res31  - For η0, the value obtained for the lower filled compound
 
 Res32  above
- For a, the value obtained for the lower filled compound
above
330 Filled Polymers

 0.026 
Φ m2
D3  31.734  A: = A = 1.201
 0.526  ( Φ m − Φ )2

call G (γ°,D) : model equation and partial derivatives

Res4: = GenFit(γ 1°3,η3,D3 ,G) : calling the nonlinear regression algorithm



n− 1
 Res4 Res42 Res4 
η3f p : =  0
+ ( Res41 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ s° p )  2 Fitting equation
 γ s° p   


 n− 1
 
Res40

r32 : = corr 
  γ 13
+ ( Res4 1 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ 13 )Res42  Res42  , η3 
     Correlation coefficient r2
  

PS + 24.0% TiO2
1.103
Shear viscosity, kPa.s

100  0.939 
10 Res4 =  54.889 
1  0.386 
0.1
1.10–3 0.01 0.1 1 10 100 1.103 1.104
Shear rate, 1/s r32 = 1

Assembling results

R 4 := A0 ← Φ
 0.056  Φ
A1 ← Res 41  54.889 
  η0
for i ∈1.. 2
 1.183  λ0
A1 + i ← Res 1i  
R 4 =  0.277  n
A4 ← Res 42  0. 386  a
A5 ← Res 40  
 0.939  σc
A6 ← r 32  1 
  r2
A
Polymers and White Fillers 331

PS + 38.6% TiO2 ΦTiO  = 0.087 Φ4: = 0.087


2

λ0: = Res11 n: = Res12 a: = Res13 Φ: = Φ4

 <= guest parameters:
 Res40  -F  or σc, the value obtained for the lower filled compound

D : =  Res41   above
4 - For η0, the value obtained for the lower filled compound
 Res42 
above
- For a, the value obtained for the lower filled compound
above

 0.939 
Φ m2
D4 =  54.889  A: = A = 1.339
 0.386  ( Φ m − Φ )2

call G(γ°,D) : model equation and partial derivatives

Re s5: = GenFit ( γ 1° 4, η4, D4 ,G ) : calling the nonlinear regression algorithm


n− 1
 Res5 Res52 Res5 
η4fp : =  0
+ ( Res51 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ s° p )  2 Fitting equation
 γ s° p   


 n− 1
 
Res50
 + ( Res51 ⋅ A ) ⋅ 1 + ( λ 0⋅ A ⋅ γ 1° 4 ) 2  Res52 , η4    Correlation coefficient r2
Res5
r42 : = corr 
  γ 1° 4    
 

1.104
1.103
η4  11.514 
Res5=  58.428 
100
η4f 10
1  0.316 
0.1
1.10–3 0.01 0.1 1 10 100 1.103 1.104 r42 = 1
γ1°4,γs°
332 Filled Polymers

Assembling results

R 5 : = A0 ← Φ
 0.087  Φ
A1 ← Res 51  58.428 
  η0
for i ∈1.. 2
 1.183  λ0
A1 + i ← Res 1i  
R 5 =  0.277  n
A4 ← Res 52  0.316  a
A5 ← Res 50  
 11.514  σc
A6 ← r 42  1 
  r2
A

A6.4.10 Assembling and Analyzing all Results


AR: = augment(augment(augment(augment(R1, R2), R3), R4), R5)

 0 0.012 0.034 0.056 0.087  Φ


16.712 19.469 31.734 54.889 58.428  η0

 1.183 1.183 1.183 1.183 1.183  λ0
 
AR =  0.277 0.277 0.277 0.277 0.277  n
 0.77 0.772 0.526 0.386 0.3166  a
 
 0 3.285 ⋅ 10−3 0.026 0.939 11.514  σc
 0.998 0.997 0.999 1 1  r2

Drawing fit curves:


i: = 0 .. 4

AR 3,i − 1

AR 0,i     
−2 −2 AR 4,i AR 4,i
AR 5,i    AR 0,i  
η(γs°,i): = +  AR 1,i ⋅  1 −  ⋅ 1 +  AR 2,i ⋅  1 − ⋅ γs°  
γs°  
Φ m      Φm   
 
 
Polymers and White Fillers 333

PS + TiO2 : Model curves


1.104

Shear Viscosity, kPa.s


1.103
100
10
1
0.1
1.10–3 0.01 0.1 1 10 100 1.103 1.104
Shear rate, 1/s
virgin PS 4.8% TiO2 14.3 %
18.0% 38.6 %

PS + TiO2 : Experimental data


1.104
Shear viscosity, kPa.s

1.103
100
10
1
0.1 –3
1.10 0.01 0.1 1 10 100 1.103
Shear rate, 1/s
virgin PS 4.8% TiO2 14.3%
18.0% 38.6%

Effect of filler level on Effect of filler level on limit


yield stress σc: viscosity η0:

15 100

10
Limit viscosity
Yield stress

50
5

0 0
0 0.05 0.1 0 0.05 0.1
Filler volume fraction Filler volume fraction

334 Filled Polymers

Effect of filler level on Yasuda parameter a:

1
Parameter a

The variations of σc, η0, and


a with the filler volume
fraction are in agreement
with the physics of the
–1
0 0.05 0.1 model.
Filler volume fraction

1.104

1.103
Shear viscosity, kPa.s

100

10

0.1
1.10–3 0.01 0.1 1 10 100 1.103 1.104
Shear rate, 1/s
virgin PS virgin PS fit
38.6% TiO2 38.6% TiO2 fit

Even for the highest TiO2 loading, for which experimental data are far to
cover a sufficient shear rate range, the fitting strategy based on lower loaded
compounds gives a reasonable curve, in agreement with the physics of the
model.
Polymers and White Fillers 335

A6.5  Expanding the Krieger–Dougherty Relationship

 − Φ ⋅B : B “intrinsic viscosity,” should be equal


η(Φ)  Φ  m to 2.5 at very low volume fraction, in
= η r (Φ) =  1 −
η(0)  Φ m  order to match the Einstein equation
: Φm maximum packing fraction

Uniform spheres, cubic packing: Φm = 0.524


Uniform spheres, random packing: Φm = 0.621
− Φ ⋅η
  Φ  m Uniform spheres, hexagonal packing:
Expanding  1 −  : Φ  = 0.741
 Φm m
Fibers, perfectly aligned, max packing:
Φm = 0.907
Prolate ellipsoids, random packing: Φm = 0.74

η r (Φ) = 1 + B ⋅ Φ + B ⋅
( Φ m⋅ B + 1) ⋅ Φ 2 + B ⋅ ( Φ m⋅ B + 1) ⋅ ( Φ m⋅ B + 2 ) ⋅ Φ 3 …
2 !⋅ Φ m 3 !⋅ Φ m 2

+B ⋅
( Φ m⋅ B + 1) ⋅ ( Φ m⋅ B + 2 ) ⋅ ( Φ m⋅ B + 3) ⋅ Φ 4 …
4 !⋅ Φ m 3

+B ⋅
( Φ m⋅ B + 1) ⋅ ( Φ m⋅ B + 2 ) ⋅ ( Φ m⋅ B + 3) ⋅ ( Φ m⋅ B + 4) ⋅ Φ 5 …
5 !⋅ Φ m 4

+B ⋅
( Φ m⋅ B + 1) ⋅ ( Φ m⋅ B + 2 ) ⋅ ( Φ m⋅ B + 3) ⋅ ( Φ m⋅ B + 4) ⋅ ( Φ m⋅ B + 5) ⋅ Φ 6 + ...
6 !⋅ Φ m 5

 1
  2 


∏ (Φ m⋅ B + i)  2



( Φ m⋅ B + i )
 3
or: η r( Φ ) = 1 + B ⋅ Φ + B ⋅  i =1
⋅Φ + B⋅  i = 1 ⋅Φ …
2!Φ m 3! Φ m 2

 3   4 


∏ ( Φ m⋅ B + i )
 4



( Φ m⋅ B + i )
 5
+B ⋅  i = 1 ⋅Φ + B⋅  i = 1 ⋅Φ +…
4!Φ m 3 5!Φ m 4

 a−1 
n  ∏
( Φ m⋅ B + i )
or abridged form: η r (Φ) = 1 + B ⋅ Φ +
a= 2
∑ 
B⋅  i = 1
a ! Φ m
a−1
 a
⋅Φ
336 Filled Polymers

Numerical illustration

Φm: = 0.74 B: = 2.5 Φ: = 0, 0.01...Φm

 a−1

n


∏(Φ m⋅ B + i )
 a : polynomial equation
ηr (Φ,n) := 1 + B ⋅ Φ + ∑
a= 2
B⋅ 
i=1

a! Φ m a−1
⋅Φ (expanded Krieger–
Dougherty)

− Φ m⋅B
 Φ 
ηrKD(Φ ) : =  1 − : Krieger–Dougherty equation
 Φ m 

ηGG( Φ ) : = 1 + 2.5 ⋅ Φ + 14.1 ⋅ Φ 2 : Guth and Gold equation

Φ: = 0.5
50
ηrKD(Φ) = 8.029
ηr(Φ, 6) = 6.549
40
ηr(Φ, 10) = 7.618
ηr(Φ, 20) = 8.017
Relative viscosity

30 ηr(Φ, 30) = 8.029
 <= a 31 terms poly-
nomial nearly per-
20
fectly macthes the
Krieger–Dougherty
10 equation

0
0 0.25 0.5 0.75
Filler volume fraction
Polynomial, 6 terms
Polynomial, 12 terms
Krieger–Dougherty
Guth & Gold
Polymers and White Fillers 337

Effect of type of packing and maximum packing fraction

Φ: = 0, 0.01...Φ m η: = 2.5

 a−1 
− Φ m⋅B n  ( Φ m⋅ B + i ) ∏

 Φ    a
η rKD ( Φ, Φ m ) : =  1 − η r ( Φ, Φ m, n) : = 1 + B ⋅ Φ + B⋅  i = 1 ⋅Φ
 Φ m  a = 2
a ! Φ m
a−1

Uniform spheres, cubic packing Uniform spheres, random packing


30 30

Relative viscosity 20
20
Relative viscosity

10 10

0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
Filler volume fraction
Filler volume fraction
Polynomial, 3 terms
Polynomial, 3 terms
Polynomial, 6 terms Polynomial, 6 terms
Krieger–Dougherty Krieger–Dougherty
   
Fibers, aligned, max packing
30 Prolate ellipsoids, random packing
30
Relative viscosity

20
20
Relative viscosity

10
10

0 0
0 0.2 0.4 0.6 0 0.2 0.4 0.6
Filler volume fraction Filler volume fraction
Polynomial, 3 terms Polynomial, 3 terms
Polynomial, 6 terms Polynomial, 6 terms
Krieger–Dougherty Krieger–Dougherty
   
7
Polymers and Short Fibers

7.1  Generalities
Short fibers and polymers are used to prepare composites essentially with
respect to the large differences between certain mechanical properties of
both. For instance, most (unfilled) vulcanized elastomers have tensile (Young)
modulus and ultimate strength in the 20–50 kPa and 2–5 MPa ranges, respec-
tively and most thermoplastics exhibit the same properties in the 1–2 GPa and
30–80 MPa ranges, respectively. Certain synthetic fibers have the same prop-
erties up to 500 GPa (modulus) and 5 GPa (strength) respectively. It means
that fibers and elastomers tensile PRoperties differ by factors of 103–107, and
fibers and plastics by factors of 102–104. Such huge differences explain why
fibers reinforcing effects are always spectacular in polymers, depending
however on fiber–matrix interfacial phenomena, fibers orientation and pro-
cessing difficulties that must be mastered for optimal results. Controlling
the orientation of short fibers in fabricated composite parts is a serious prob-
lem in most processing operations, while certain processes are more prone
to improvement than others, depending on the complexity of the associated
flow fields. At best, what can be achieved with short fiber-filled polymer sys-
tems is either random or preferred orientation, never perfect orientation. It
follows that the actual reinforcing effects obtained when adding short fibers
to a polymer can be considered as being bounded by two extreme (ideal) situ-
ations: either all fibers are perfectly aligned with respect to the main strain
axis, or they are all perpendicular. Figure 7.1 illustrates this concept with an
ideal system made of perfectly aligned (long) fibers in a matrix. The tensile
moduli of the fibers and the matrix are Efib and Emat, respectively.
If the fibers are all aligned in the direction of the strain, both the matrix
and the fibers experience the same, uniform strain, and the composite modu-
lus is the upper-bound, calculated as follows:

Eupper = Efib × Φ fib + Emat × ( 1 − Φ fib ) (7.1)

where Φfib is the volume fraction of fibers. This equation corresponds to the
simple mixing rule and is sometimes referred to, in mechanics, as the Voigt
339
340 Filled Polymers

Matrix with
modulus Emat

Fibers with
modulus Efib

Strain parallel Strain perpendicular


to fibers to fibers

80
Efib = 73 GPa

60
Modulus, GPa

Eupper
40

20 Elower

Emat = 3 GPa
00 0.2 0.4 0.6 0.8 1
Fibers volumic fraction Φfib

Figure 7.1
Upper and lower bounds in (long) fiber composites; the upper and lower bound moduli curves
where calculated with the typical tensile modulus values for E-glass fibers (Efib = 73 GPa) and
for polyamide 6 (Emat = 3 GPa).

average, with respect to the assumption made by this author that “in a multi-
phase body, the average strain of each phase is equal to the applied strain.”1
If the strain is applied perpendicularly to the fibers orientation, both the
fibers and the matrix experience the same, uniform stress, and the composite
modulus is the lower bound, i.e.:

1 Efib Emat
Elower = = (7.2)
Φ fib 1 − Φ fib Emat Φ fib + Efib ( 1 − Φ fib )
+
Efib Emat

This equation corresponds to the harmonic mixing rule (or inversed mix-
ing rule) and is also referred to as the Reuss average, with respect to the
assumption made by this author that (in a multiphase body) “the average
stress in each phase is equal to the applied stress.2”
Polymers and Short Fibers 341

The above equations correspond to the intuitive understanding that in the


former case, the mechanical response of the laminar composite is essentially
dominated by the performance of the fibers, thus giving the upper-bound
reinforcing effect, while in the latter case, the matrix (and the fibers-to-matrix
adhesion) is playing the key role and dictates the lower-bound reinforcement.
As we will see, some micro-mechanical models for short fibers composites
clearly emphasize such upper and lower limits, and it follows that what can
really be achieved when manufacturing a short fiber-reinforced polymer
object is indeed between those two extremes.
In terms of scientific research, a large variety of fiber-polymer systems
have been (and are still) studied, with new compositions constantly investi-
gated. However, in terms of industrial realizations, the list of useful systems
is reduced to a few tens, as described in Table 7.1.
As one might expect, any fiber-polymer system is unique, not only because
of the differences in the respective properties of the components, but also
because the fiber surface properties and the polymer chemistry have impor-
tant effects. Not all aspects are yet understood and any generalization would
indeed be abusive. A clear demonstration of this fact is offered when com-
paring the effect of a given type of short glass fiber (E-glass), at constant
loading (30%) in different thermoplastics (Table 7.2).
As can be seen, the effect of the filler is strongly depending on both the poly-
mer and the property considered. Generally, the ultimate tensile and flexural
strengths increase upon addition of short glass fibers (SGF), while the elongation
at break decreases. The effect of fibers on the impact resistance depends essen-
tially on the semicrystalline or amorphous nature of the polymer (see Figure 7.2).
A net increase is observed with the former, a loss with the latter. The loss is par-
ticularly impressive in the cases of polycarbonate and poly(phenylene oxide);
which explains why these polymers are generally used alone in applications
where impact resistance is essential, but rarely as components of composites.
SGF bring a benefit on heat distortion temperature, but generally larger
with semicrystalline polymers (Figure 7.3). Whatever is the polymer, fibers
reduce the thermal dilation, which is one of the reasons in using fibers-filled
­composites in applications where tight dimensional tolerance of parts is

Table 7.1
Most Frequent Polymer-Short Fiber Systems for Industrial Applications
Polymers Short Fibers
Polypropylene PP Synthetic:
Polyamides PA Short glass fibers, SGF
Polyvinylchloride PVC Aramids (staple, pulp), SAF
Saturated polyesters PET, PBT Chopped carbon fibers, SCF
Polyphenylene oxide PPO Natural:
Polycarbonate PC Various cellulose fibers, untreated or treated
Vulcanizable elastomers   Natural, e.g., cellulose
NR, SBR, BR   Synthetic, e.g., aramids
342

Table 7.2
Effects of 30%wt Short E-Glass Fiber in Different Thermoplastics
PA-66 PP PC PBT PPO
Gain/Loss Gain/Loss Gain/Loss Gain/Loss Gain/Loss
Property Unit Pure in Cpd Pure in Cpd Pure in Cpd Pure in Cpd Pure in Cpd
Ultimate tensile strength MPa 77 +98 30 +53 60 +62 55 +78 55 +75
Elongation at break % 150 –146 300 –297 110 –106 150 –147 50 –46
Ultimate flexural strength MPa 100 +145 40 +70 87 +93 90 +90 89 +59
Izod impact, notched J/cm 179 +250 125 +179 2500 −2143 179 +89 893 –536
HDT (1.8 MPa) °C 71 +182 70 +79 139 +7 69 +144 120 +23
Thermal dilatation °C–1 50 –33 90 –60 68 –46 75 –50 50 –30
coefficient ( × 106)
Filled Polymers
Polymers and Short Fibers 343

Impact resistance Pure polymer


3000
Polymer + 30% GF
IZOD impact, notched, g/cm
2500

2000

1500

1000

500

0
PA-66 PP PC PBT PPO

Figure 7.2
Effect of 30% short E-glass fibers on the impact resistance of various thermoplastics.

Heat resistance Pure polymer


300
Polymer + 30% GF
250
HDT (1.8 MPa),°C

200

150

100

50

0
PA-66 PP PC PBT PPO
Figure 7.3
Effect of 30% short E-glass fibers on the heat distortion temperature of various thermoplastics.

an issue (e.g., connectors for electronic and various applications in electro


mechanics).
The obvious difference between particulate (e.g., carbon black, silica, and
fine minerals) and fibrous fillers is the large aspect ratio of the latter, which
makes them particularly sensitive to strong (i.e., orienting) flow in process-
ing operations. In addition, one understands easily that most of the short
fibers-related benefits described above also strongly depend on the average
orientation of the fibers. If, for a given composition, a majority of the fibers
are aligned in the stretching direction, then the tensile properties must nec-
essarily be close to the upper bound limit. Conversely, for better flexural
properties, most of the fibers must be oriented perpendicularly to the load
application. Such remarks highlight the importance of mechanical models for
fibers-filled composites, not only with respect to their expectedly predictive
capabilities, but also as convenient tools for understanding the very origin of
the benefits obtained, as well as the likely sources of failure or problems.
344 Filled Polymers

7.2 Micromechanic Models for Short Fibers-Filled


Polymer Composites
7.2.1  Minimum Fiber Length
Short-fibers reinforced composites have been developed, used and analyzed for
long as follow-up of many successful engineering developments, most of them
on a rather pragmatic basis (trial-and-error). In parallel, there as been a continu-
ing effort to develop theoretical expressions that would allow the mechanical
properties of a composite to be predicted, knowing only the properties of the
polymer matrix and the fibers. Let us consider a homogeneous matrix in which
short anisometrical particles (e.g., ellipsoids, rods, fibers) have been randomly
dispersed. The rigidity of the dispersed inclusions is considerably larger than the
one of the matrix. Then let us submit this composite material to a strain (what-
ever is the mode of deformation). It is commonsense that stress singularities will
developed in the matrix neighboring regions of the particles. To calculate the
overall nonuniform stress field of the composite would obviously be a formida-
ble task, however not really necessary because the problem can be tackled either
in terms of average stress and strain, or by considering the local situation and
extending it to the whole composite. Such considerations are the essence of so-
called micromechanical approaches. There are a many excellent reviews3,4 and a
few textbooks5–7 dealing with micromechanic models whose objective is to pre-
dict the average elastic properties of composite materials. The following ­section
will be limited to a short description of the most easy-to-handle models (in the
author’s opinion and experience), leaving purposely aside those approaches that
need extensive numerical simulation efforts to be applied.
Nearly all micromechanics models for short-fibers filled systems consider
the same basic assumptions:

• Both the fibers and the matrix are linearly elastic; the matrix is iso-
tropic and the fibers have constant properties along their length.
• The fibers are axisymmetric with a narrow distribution of shape and
size, so that they can be idealized as rods, essentially characterized
by an (average) aspect ratio, i.e., the length-to-diameter ratio L/D.
• There is perfect bonding between the fibers and the matrix, and it
remains so during deformation of the composite. No such effects as
interfacial slip, fiber-matrix decohesion or matrix microcracking are
considered.
• Fibers concentration is finite but not so large to have direct contact
between them.

In short-fibers composites, loads are transferred from the matrix to the


fibers in a zone near the fiber end, so that how the stress is distributed around
a fiber is an important aspect. A basic approach to this problem is the classical
Polymers and Short Fibers 345

“shear-lag analysis” by Rosen and Dow.8 Fiber-end geometry and adjacent


fibers effects are ignored, so that the fiber stress is zero at the end and
increases gradually as the load is transferred from the matrix to the fiber. The
maximum reinforcing effect is then achieved when fibers are long enough for
complete load transfer to occur. If both fibers and matrix are elastic bodies,
there is a minimum fiber length Lmin for this optimal load transfer, i.e.

1 Efib 1 − Φ fib
Lmin = D (7.3)
2 Gmat Φ fib

where D is the fiber diameter, Efib and Gmat the fiber elastic and the matrix shear
moduli respectively, and Φfib the volume fraction of fibers. The minimum fiber
aspect ratio Lmin/D decreases thus slightly with the increasing fiber loading but
is much depending on the fiber-to-matrix modulus ratio. As Efib/Gmat increases
the minimum fiber aspect ratio becomes very large, for instance with Efib = 73
GPa and Gmat = 73 kPa, one gets Lmin/D = 786 with D = 0.5 μm (see Appendix 7.1
for a numerical illustration of Equation 7.3). But the shear lag analysis suffers
in fact from a number of excessive hypotheses and, for instance, finite element
analysis has proved that there is a strong interfacial stress concentration at fiber
ends, which contributes of course to composite’s hysteresis. The demonstration
provided by Equation 7.3 that there is a minimum fiber aspect ratio required for
effective stress transfer remains however fully valid and is of importance with
respect to the processing of fibers-filled composites. Indeed, fibers inevitably
brake during processing operations, and the lower the fiber ratio, the lower the
reinforcement. However, because the Efib/Gmat ratio is generally large, substan-
tial reinforcement can still be obtained, even with reduced L/D ratio.

7.2.2  Halpin–Tsai Equations


The Halpin–Tsai equations are a set of empirical relationships that have long
been used to predict the mechanical properties of composite materials.9,10
They express the composite property with respect to those of the matrix and
of the reinforcing material, their respective volume fraction and the filler
geometry. Essentially, these equations give the (mechanical) property of a
composite Pcpd in terms of the corresponding property of the matrix Pmat and
the reinforcing phase Pfil through the following relationships:

 Pfil 
 1 + ζ µ Φ fil   P  − 1
Pcpd = Pmat     with   µ =
mat
(7.4)
 1 − µ Φ fil   Pfil 
 P  + ζ
mat

The Halpin–Tsai equations were originally developed for continuous-fiber


composites with respect to early self-consistent models for ideal systems
346 Filled Polymers

with perfectly aligned fibers. The mechanical properties that can typically
be predicted using the above equations are the longitudinal and transverse
tensile moduli, the shear modulus and the Poisson’s coefficient. Depending
on the property and the geometry of the filler particles, different expressions
must be used for ζ, as summarized in Table 7.3.
The above equations are essentially empirical (i.e., the ζ term has a weak
theoretical background) but their validity is confirmed by numerous exper-
imental measurements, at least for moderate filler volume fractions. There
were also found in excellent agreement with finite element calculations on
idealized short fibers composites.4 For filler fractions larger than 0.4, Hewitt
and de Malherbe11 have proposed to make the ζ term depending on the filler
volume fraction by adding a component equal to 40 Φ10 (found by curve fit-
ting). As demonstrated through calculation in Appendix 7.2, this extra term
brings indeed negligible changes until Φ is larger than 0.4. Halpin and
Kardos10 noted that ζ must lie between 0 and infinity. If ζ = 0, Equation 7.4
reduces to the harmonic mixing rule (Equation 7.2), and if ζ = ∞, the Halpin-
Tsai form becomes the simple mixing rule (Equation 7.1). Choosing the appro-
priate value for ζ is in fact the most critical aspect in using the Halpai-Tsai
equations, because the relationships quoted in Table 7.3 have been obtained
by considering somewhat idealized systems, either with respect to the orien-
tation of the filler (relevant for fibers, plates, and whiskers only) and/or with

Table 7.3
Detailed Expressions for the ζ Parameter in Halpin–Tsai Equations
Depending on Filler Particle’s Geometry
Longitudinal Tensile Transverse Tensile Shear Modulus
Particle Geometry Modulus E11 Modulus E22 G12
Spherical particle ζ = 2 ζ = 2 ζ = 1
L
Oriented short fibers ζ =2 ζ = 2 ζ = 1
D
1.73
 L  L  L +W 
Oriented plates ζ=2  ζ=2  ζ=
T W  2 T 
1.73
 L L
Oriented whiskers ζ=2  ζ = 2 ζ= 
 D  D

Notes:
• L and D are respectively, the length and diameter of fibers and whiskers.
• The minimum effective length for short fibers can be calculated with Equation 7.3.
• L, T and W are respectively, the length, thickness and width of plates.
• Whatever the geometry of the filler particle, the Poisson’s ratio of the composite is
νfil Φ fil + (1 − Φ fil ) νmat .
• Whiskers are crystalline metallic tiny, filiform hairs that spontaneously grow from
metallic surfaces. Whiskering is seen on elemental metals and on alloys. In the pres-
ent context, whiskers must be seen as “curved” or “curled” filaments, in contrast with
short fibers which are essentially rectilinear rods.
Polymers and Short Fibers 347

the implicit assumption that all particles are identical since their geometry is
expressed through a set of single numbers. This latter assumption appears
less critical than the actual fibers’ orientation. A numerical illustration of the
mathematical virtues of the Halpin-Tsai equations is given in Appendix 7.2, as
well as typical curves for the longitudinal, transverse and shear moduli, in the
case of polypropylene-SGF composites. In short fibers-filled systems, the fiber
aspect ratio plays a role only in the equation for the longitudinal modulus.
Figure 7.4 shows typical curves as calculated with the Halpin-Tsai equations
and a comparison of the calculated longitudinal and transverse moduli with
(average) measured flexural modulus data on commercial PP-SGF systems.
As can be seen, experimental data fall in between calculated E11 and E22
moduli likely because, in the tested commercial samples, fibers are oriented
neither longitudinally nor transversally. For a complete random orientation
of fibers, the overall modulus would be given by an appropriate fractional
summation of longitudinal and transverse moduli, i.e.12

3 5
Erandom = E11 + E22 (7.5)
8 8

It follows that, practically, the average fibers’ orientation could be expressed


through an adjustable parameter X, such that Equation 7.5 can be rewritten
as follows:

Erandom (Φ) = X E11 (Φ) + (1 − X ) E22 (Φ) (7.6)

30 20
Calculated longitudinal
Tensile or flexural modulus, GPa

Calculated transverse Maximum fiber packing : 0.9


Experimental
Shear modulus, GPa

20

10

10

0 0
0 0.1 0.2 0.3 0 0.5 1
Fiber volume fraction Fiber volume fraction

Figure 7.4
Typical curves as calculated with Halpin–Tsai equations and parameters for short glass fiber–
polypropylene composites; parameters used in calculation: short glass fibers: Efib = 77.0 GPa,
νfib = 0.20, L = 1 mm; D = 5 μm, i.e., a fiber aspect ratio of 200; polypropylene, Emat = 1.14 GPa,
νmat = 0.43; experimental data are average flexural modulus data from various suppliers of
PP-SGF composites. For both the fibers and the matrix, the shear moduli were calculated using
the standard equation :G = E/[2(1 + ν)].
348 Filled Polymers

30

Longitudinal modulus

20
Modulus, GPa

4.5
X=
8

10

Transverse modulus

0
0 0.1 0.2 0.3
Fiber volume fraction

Figure 7.5
Fitting experimental data with Halpin–Tsai equations and a parameter for average short fiber
orientation; parameters used in calculation: short glass fibers: Efib = 77.0 GPa, νfib = 0.20, fiber
aspect ratio of 200; polypropylene, Emat = 1.14 GPa, νmat = 0.43.

As shown in Figure 7.5, a better fit of experimental data is obtained with the
appropriate value for the “orientation parameter” X.
It is pretty obvious that fiber orientation distribution and how to control
it during processing are amongst the most critical variables which affect
the mechanical properties and hence the efficient use of short-fiber compos-
ites. Certain processing techniques allow some control of fiber orientation,
essentially because important extensional flow fields and the associated flow
anisotropy effects can be generated. It has been shown for instance that fiber
aspect ratio, fiber-matrix interaction, processing tools geometry, shear rate,
temperature, and fiber volume loading are among the very important param-
eters which control the final fiber orientation in processes such as extrusion,
injection and transfer molding.13
It is difficult to assess the fiber orientation distribution. However E11 and
E22 are easily calculated providing the matrix and fibers parameters and the
fibers fraction are known. It follows that the average orientation of fibers in a
given composite sample can be estimated from the measured modulus. Such
an approach was indeed used by Leblanc et al.14 to study the mean fibers’
orientation in injection molded fatigue test specimens (ASTM D1708) with
commercial short glass fiber composites with either polybutylene therephta-
late (PBT) or copolyamide 6/6T (PA/PAT) as matrix material. Parallelepiped
samples (3 × 5 × 10 mm) were cut out of the fatigue specimens, parallel to the
cavity filling axis, and tensile Young modulus was measured (ASTM D638).
Average short glass fibers dimensions were measured by electron micros-
copy and found to be L ≈ 300 µm and D ≈ 10 µm. Glass fibers were of the
E-glass type (i.e., E11 = 73 GPa). The longitudinal and transverse moduli were
Polymers and Short Fibers 349

calculated using the Halpin–Tsai equations, then the orientation parameter


was calculated using (see details in Appendix 7.2):

Emeas − E22
X= (7.6a)
E11 − E22

For randomly aligned fibers, X would be equal to 0.375. As seen in Table 7.4,


all calculated X values are higher than 0.375 and clearly decrease with
increasing fiber content. It follows that, as expected, fibers in the fatigue test
specimens tend to be aligned along the longitudinal mold axis, but the higher
the fiber content, the less effective the longitudinal flow induced orientation,
probably due to fiber hindrance. Microphotographs of polished sections cut
in the fatigue specimens essentially confirmed the general but not perfect
orientation of fibers along the mold filling direction.
Even for well aligned short fibers systems however, the Halpin–Tsai equa-
tions do not compare well with experimental results for large volume frac-
tions and their theoretical maximum applicability limit is the maximum
hexagonal packing of perfectly aligned fibers, i.e., π/(2 3 ) ≈ 0.907 . As shown
in Figure 7.4, experimental data for current fiber loadings lie between the lon-
gitudinal and the transverse calculated tensile moduli. This corresponds of
course to the fact that in real short fibers systems the alignment is neither per-
fect nor completely at random, as demonstrated above. For large fibers fraction
systems (i.e., above 0.4–0.5) Nielsen and Lewis15,16 have suggested that the filler
packing limit must be taken into consideration, as an upper-bound parameter,
through the following modified form of the Halpin–Tsai equation, i.e.

 1 + ζ µ Φ fil 
Pcpd = Pmat  (7.5)
 1 − F(Φ fil ) µ Φ fil 

Table 7.4
Calculating the Average Orientation from Measured Modulus and
Halpin–Tsai Equations for PBT and PA/PAT Composites with Short
Glass Fibers
Material ΦFiber Modulus (GPa) Orientation Parameter X
Virgin PBT – 2.600 –
PBT + 20% SGF 0.1115 6.475 0.606
PBT + 30% SGF 0.1765 8.445 0.587
PBT + 50% SGF 0.3288 14.265 0.576
Virgin PA/PAT – 3.200 –
PA/PAT + 25% SGF 0.1308 8.431 0.697
PA/PAT + 35% SGF 0.1938 10.771 0.657
PA/PAT + 50% SGF 0.2962 14.930 0.636
350 Filled Polymers

with various explicit forms for the functional F(Φfil), for instance:

 1 − Φ max 
F(Φ fil ) = 1 +   Φ fil (7.5a)
 Φ max
2

or

1   Φ Φ 
F(Φ fil ) = 1 − exp  fil max   (7.5b)
Φ fil   Φ fil − Φ max 

The functional F(Φfil) must be such that the product F(Φfil) × Φfil fulfills the
following boundary conditions:

d [ F(Φ fil ) × Φ fil ]


F(Φfil) × Φfil = 0   at   Φfil = 0, = 1   at   Φfil = 0
d Φ fil

and   F(Φfil) × Φfil = 1   at   Φfil = Φmax.

The two first conditions are imposed by the fact that as the filler fraction
goes to zero, one must recover the Einstein equation. Equations 7.5a and 7.5b
were selected by Lewis and Nielsen as amongst the simplest ones fulfilling
such conditions, but otherwise have no theoretical justification. Equation 7.5a
is the equation of a straight line and Equation 7.5b has a maximum above
0.75 × Φfil. When multiplied by Φfil, both functions nearly superimpose up
to Φfil = 0.12 and exhibit minor differences above this level. In fact, as shown
in Appendix 7.3 (Section A7.3.2), such modifications of the Halpin-Tsai equa-
tions do not bring much changes in the fiber fraction range of practical inter-
est (0–0.25).
For short-fiber composites, it is clear that the above equations are expected
to give good predictions only if the fibers are somewhat aligned. As the
fiber aspect ratio L/D increases, the longitudinal modulus of the composite
increases and tends to the limit Efib × Φfib of long-fiber composites. As we
have seen with Equation 7.3, the fiber aspect ratio must be larger with higher
Efib/Gmat ratio. It follows that reinforcement with short fibers is more efficient
for relatively small Efib/Gmat ratio. For a given polymer matrix, with respect
to the range of practical fiber aspect ratio (e.g., 200–300), at loadings that are
compatible with processing constrains and requirements, the longitudinal
tensile modulus of composites reach an asymptotic limit, and using short
fibers with larger tensile modulus brings only minor changes. The trans-
verse and the shear moduli however are independent of the fiber aspect ratio,
and if the fiber modulus is large compared to the matrix modulus, Ey and G
Polymers and Short Fibers 351

become also independent of the fiber modulus, in which case, the following
simple relationships can be considered:

 1 + 2 Φ fib 
Ey = Emat   (7.6a)
 1 − Φ fib 

 1 + Φ fib 
Gcpd = Gmat   (7.6b)
 1 − Φ fib 

7.2.3  Mori–Tanaka’s Averaging Hypothesis and Derived Models


Mori and Tanaka17 considered likely interactions between inhomogeneities
in complex systems (explicitly metallic alloys in their paper) and introduced
the concept of an average field that would include the inhomogeneities and
the surrounding matrix. Essentially the Mori–Tanaka hypothesis can be
stated as follows: within a heterogeneous concentrated composite submitted to a
strain, each particle “feels” a far-field strain that is equal to the average strain in the
matrix. Alternatively this average stress/strain concept could be defined as
follows: when submitted to a given applied stress, the average stress in the
matrix differ from the applied one due to the presence of inclusions; how-
ever, the (volume) average of the affected parts in both the inclusions and the
surrounding matrix must vanish in order to satisfy the equilibrium condi-
tion. As described in their paper, such a hypothesis allows to calculate the
average internal stress (upon strain) in a matrix containing inclusions. The
approach is not easy and consists of complex mathematical manipulations
of field variables with respect to the concept of equivalent inclusions. The
Mori and Tanaka hypothesis was however considered by several authors,
namely Benveniste18 who provided a much simpler description of the aver-
age strain concept. Benveniste’s explanation prompted several authors to for-
mulate simple expressions for moduli of materials, when some simplifying
assumptions can be made. Otherwise numerical methods are needed to use
the so-called Mori–Tanaka model.4
Tandon and Weng19 have derived a complete set of explicit relationships
for moduli of a composite model in which randomly distributed ellipsoi-
dal particles (i.e., short-fiber like) are unidirectionally aligned. The Mori–
Tanaka’s average strain concept is also used as the transformation tensor
described by Eshelby20 when he solved the problem of an ellipsoidal inclu-
sion in an elastic field. The Eshelby’s transformation tensor is a 4th order
tensor whose components depend only the aspect ratio of the inclusion and
the elastic moduli of the matrix. Some of the equations obtained by Tandon
and Weng must however be solved iteratively and handling them implies
calculating at first quite a impressive number of “constants,” in fact various
352 Filled Polymers

combinations of matrix and inclusions moduli and Poisson’s ratio, and


the appropriate nonzero components of the Eshelby’s tensor with respect
to the case considered. Explicit formulas were given in the Tandom and
Weng’s paper and are reproduced in Appendix 7.4 (Sections A7.4.1 through
A7.4.3), with a numerical illustration of their virtues in predicting the ten-
sile and the shear moduli of short-fibers filled composites. With respect to
fiber volume fraction, all the calculated moduli fall between the lower and
upper bounds (Equations 7.1 and 7.2) and are relatively insensitive to the
fiber aspect ratio, except the longitudinal (tensile) modulus, (see Figure 7.6).
As the aspect ratio decreases, the modeled E11 becomes close to the lower
bound prediction, but when the L/D ratio is larger than 100, no difference
is seen between the Mori–Tanaka’s approach and the upper bound predic-
tion. As shown in Figure 7.6, one can hardly see a difference in the modeled
shear modulus when the fiber aspect ratio varies from two to several hun-
dreds. As the predicted shear modulus remains close to the lower bound
whatever the fiber volume fraction, it is quite clear that the average strain
approach is far to correspond to common experimental observations which
reveal a steeper variation of the shear modulus when the short fiber content
increases.
If the fiber aspect ratio is set to one (i.e., spherical particles), one notes that
there is a mathematical singularity in the Eshelby tensor and the Tandon and
Weng explicit equations for short fibers cannot be used. However, in such
a case, there are some simplifications in the tensor and particularly simple

100 40

L = 200 30
Tensile modulus, GPa

Shear modulus, GPa

50 20
20

10 L
2 = 200, 20, 2
D

0 0
0 0.5 1 0 0.5 1
Fiber volume fraction Φ Fiber volume fraction Φ
Emat = 1.19 GPa νmat = 0.35 Gmat = 0.44 GPa Upper bound modulus
Efib = 77.0 GPa νfib = 0.20 Gfib = 32.1 GPa Lower bound modulus

Figure 7.6
Mori-Tanaka’s average strain approach; sensitivity to fiber aspect ratio.
Polymers and Short Fibers 353

expressions are obtained (see Appendix 7.4, Section A7.4.9), for instance for
the shear modulus:

 
  Φ fill  ( Gfill − Gmat ) 
Gcpd = Gm 1 +   (7.7)
  ( 1 − Φ fill )   2 ( 4 − 5 νmat )  
  15 (1 − ν )  fill( G − G mat ) + Gmat 
 mat 

where Gcpd, Gmat and Gfill are the moduli of the composite, the matrix and the
filler respectively, and νmat the Poisson’s ratio of the matrix.
In the particular case of polymer films containing “liquid fillers,” Gao and
Tsou21 have extended the Tandon and Weng’s work to derive easy-to-handle
expressions for the moduli and the Poisson’s ratio. Through a comparison
with finite element calculations on idealized arrays of fibers, Tucker and
Liang4 have evaluated models derived from the Mori–Tanaka’s average field
concept, the (easier-to-handle) Halpin–Tsai equations and other models. They
found that the Halpin–Tsai equations give reasonable estimation for stiffness
but the best predictions (of finite element calculations) were obtained with
the Mori–Tanaka model.
The evaluation by Tucker and Liang is however somewhat artificial since
they considered (for their Finite Element (FE) calculations) an ideal short
fibers composite, with a model array (i.e., arrangement) of the fibers. A more
pertinent evaluation is made with respect to measured data (see Appendix
7.4, Section A7.4.8 for details). Figure 7.7 compares the prediction of the
Mori–Tanaka’s average strain approach with experimental data on commer-
cial short glass fibers-filled polypropylene composites. The following data
were used in calculating the model curves: short glass fibers, Efib = 77.0 GPa,
νfib = 0.20, L = 1 mm; D = 50 μm, i.e., a fiber aspect ratio of 20; polypropylene,
Emat = 1.14 GPa, νmat = 0.43. For both the fibers and the matrix, the shear modu-
lus was calculated using the standard equation: G = E/2(1 + ν). Essentially the
E11 longitudinal modulus was calculated, to be compared with both the flex-
ural and tensile moduli (which are generally equal for isotropic systems and
appear to be also equivalent for filled thermoplastics; see Figure 7.8). For the
sake of comparison, the much simpler modified Guth and Golf equation, i.e.,
Ecpd = Emat (1 + 2.5f Φ + 14.1f 2 Φ2) with f an (empirical) anisometry factor, was
used to draw other model curves. As can be seen the Mori–Tanaka’s approach
only allows experimental data to be met at low fiber volume fraction. Above
Φ ≈ 0.10, experimental moduli are systematically higher than the model and
the simpler modified Guth–Gold equation with f = 4.2 gives a better fit.

7.2.4  Shear Lag Models


So-called shear lag models are really micromechanical by nature as they
consist in considering the behavior of the matrix near the fiber surface. Such
354 Filled Polymers

20 20
Flexural modulus, GPa

Tensile modulus, GPa


10 10

0 0
0 0.05 0.10 0.15 0.20 0.25 0 0.05 0.10 0.15 0.20 0.25
Fiber volume fraction Φ Fiber volume fraction Φ
Average data, various suppliers Mori–Tanaka's averaged strain
Single supplier's data Modified Guth & Gold

Figure 7.7
Comparing the Mori–Tanaka’s average strain model with experimental data on commercial
short glass fiber filled polypropylene composites.

20
5
18 4

16 3
2
Flexural modulus, GPa

14
1
12 0
0 1 2 3 4 5 Pure polymers [D.W. Van Krevelen
Properties of Polymers,
10 3rd Ed., Elsevier (2003)]
8 PP + CaCO3 [Table 6.7]
PP + Talc [Table 6.8]
6 PP + Mica [Table 6.9]

4 PP + SGF [Table 7.6]


PA6 + SGF [Table 7.7]
2
PA11 + SGF [Table 7.9]
0
0 5 10 15 20
Tensile modulus, GPa

Figure 7.8
Comparing flexural and tensile moduli for thermoplastic polymer systems.
Polymers and Short Fibers 355

models are algebraically simple and have an obvious ­physical ­significance.


The shear lag analysis made by Cox22 focuses on a single fiber of length L
and radius R, embedded in a concentric cylindrical shell of matrix with
radius r; see Figure 7.9 (note: the matrix shell thickness is thus r–R).
Essentially, the model intends to predict the longitudinal (tensile) modulus
E11, so only the axial stress and strain are of interest. Poisson’s ratios effects
are neglected and it is considered that when the composite is submitted to a
stress, there is a difference in displacement between the fiber surface and the
outer surface of the cylindrical matrix shell. The key hypothesis of the shear
lag model is that the axial shear stress at the fiber surface is proportional to
this difference. With the boundary condition of zero stress at fiber’s ends,
an expression for the average fiber stress is obtained, which combined with
a corresponding average fiber strain through a so-called “efficiency factor,”
allows to derive an equation for the longitudinal modulus. The model is then
completed by combining the average fiber stress with an average matrix stress
in order to obtain a modified rule of mixture for the axial modulus, i.e.

E11 = nL Φ fib Efib + (1 − Φ fib ) Emat (7.8)

  βL 
 tanh  2  
where nL is the “efficiency factor,” i.e., nL = 1 −  (7.8a)
 βL 
 2 

2 Gmat
β= (7.8b)
 r
R Efib ln  
2
 R

2
1

Matrix shell
3
Fiber
Hexagonal fiber packing
150
Mesophase thickness

π
100 2 3
micron

R
r L
50

0
0 0.2 0.4 0.6 0.8 1
Fiber volume fraction

Figure 7.9
Shear lag model.
356 Filled Polymers

Efib and Emat are the tensile moduli of the fiber and the matrix respectively,
Gmat the shear modulus of the matrix, L and R the length and radius of the
fiber. In using the above equations for a real composites, i.e., many fibers (all
perfectly aligned) with average length L and radius R, the remaining prob-
lem is to have a reasonable estimation for the matrix shell radius r. It is quite
obvious that the shell thickness must decrease with increasing fiber volume
fraction (see the inset in Figure 7.6). Several choices are possible for r, all
depending on the fibers spatial arrangement, and are conveniently expressed
through the following equality:

Kr
r=R (7.8c)
Φ fib

Kr is a constant (in fact the packing ratio of the fibers in the composite) that
depends on the assumption made about the fibers packing mode. Cox consid-
ered an hexagonal packing and chose r as the distance between the centers of
two neighbor fibers (see Table 7.5). This choice is not really realistic however
since, for touching fibers, r would be equal to 2 R with no more matrix layer
between the fibers. Another apparently easy choice is Kr = 1 so that the fiber
and its matrix shell have together the same volume fraction as the fibers in
the composite.23,24 However we do not consider this choice as realistic since Kr
cannot be larger than π/2 3 ≈ 0.907 , i.e., the maximum hexagonal packing
fraction for fibers of equal diameter. Other authors have considered a square
array of fibers, with r as half the distance between centers of nearest fibers.
In such as case, Kr is close to experimental values for the maximum fibers
­packing fraction. Except the Cox choice, all the others give nearly identical

Table 7.5
Shear Lag Model; Values of Kr
Fibers Arrangement Matrix Shell Radius Kr

r 2π
Cox = 3.628
3

r π
Hexagonal = 0.907
2 3

r
π
Square = 0.785
4
Polymers and Short Fibers 357

smooth decreases of the matrix shell thickness with larger fibers fraction. It is
clear however that each assumption gives a somewhat different dependence
of the efficiency factor Φfib and hence (slightly) different values for the pre-
dicted modulus. Larger values of Kr lead to lower values of E11.
Whatever is the arrangement of the fibers, the mathematical form of
Equation 7.8c dictates a smooth decrease of the shell thickness (r–R) as
the fiber content increases, with a significant difference between the Cox
arrangement and the other arrangements (hexagonal and cube), owing to
the particular choice for Kr (see upper left graph in Figure 7.6). The key fac-
tor in the shear lag model remains however the fiber aspect ratio, as eas-
ily demonstrated when comparing calculated E11 values with experimental
data on short fibers-filled composites (Figure 7.10; see calculation details in
Appendix 7.5). Depending on the fibers arrangement selected, an adequate
choice for the fiber ratio allows measured data to be well fitted, up to fiber
fractions of around 0.18, essentially because the model exhibits a slight cur-
vature with respect to Φfib. For commercial PP–SGF composites, the fibers
ratio is likely to be between 20 (hexagonal and cube arrangement) and 30
(Cox arrangement), if one considers the shear lag model with confidence.

600 30

Kr =
Matrix shell thickness,

L/D
Longitudinal/flexural

2π 3
Kr =
400 20
modulus, GPa

3 50
micron

30
Kr = π
20
200 2 3 π 10
Kr =
4 10

0 0
0 0.5 1 0 0.1 0.2 0.3
Fiber volume fraction Fiber volume fraction

30 30
π L/D L/D
Kr = Kr = π
Longitudinal/flexural
Longitudinal/flexural

4
2 3 50 50
modulus, GPa

20
modulus, GPa

20
30 30
20 20

10 10 10 10

0 0
0 0.1 0.2 0.3 0 0.1 0.2 0.3
Fiber volume fraction Fiber volume fraction

Figure 7.10
Typical features of the shear lag model; effect of fiber aspect ratio on calculated modulus,
compared with experimental data; material parameters: Short Glass Fibers, Efib = 77.0 GPa,
νfib = 0.20, L = 1 mm; D = variable; polypropylene, Emat = 1.14 GPa, νmat = 0.43.
358 Filled Polymers

When applying the above model to experimental data, one considers


implicitly that the L/D ratio is an average value, whatever is the actual dis-
tribution of fiber lengths. Would this distribution be known, one could obvi-
ously calculate a distribution of efficiency factors. Numerical handling of the
above equations with respect to fictitious fiber length distributions demon-
strates very little benefit in this approach with respect to a simple L/D ratio
average. A more significant concern is that, in the shear lag model, the fibers
are considered as perfectly aligned. In practice this is never the case and
Equation 7.8 could be rewritten as:

E11 = O f nL Φ fib Efib + (1 − Φ fib ) Emat (7.9)

where Of is a coefficient that depends of the distribution of fibers orienta-


tion. For a random 2D distribution of fibers orientation, Of = 1/3, and for a
random 3D distribution, Of = 1/6. It is fairly obvious that, with all the other
parameters being kept constant, a random orientation of the fibers gives a
lower longitudinal modulus. Through numerical simulation, some authors25
have however concluded that shear-lag type models do not apply to random
fiber networks, because most of the axial stress is transferred directly from
fiber to fiber rather than through an intermediate shear-loaded matrix layer.
It is worth noting that the shear lag model is in complete contradiction with
the average stress and strain field concepts of Mori–Tanaka’s and followers,
but is in line with the concept of mesophase introduced by Theocaris26 (see
Chapter 5, Section 5.1.9).

7.3  Thermoplastics and Short Glass Fibers


Lightweight reinforced thermoplastics materials with great strength and
stiffness are needed in a number of so-called technical applications. Therefore
short glass fibers are used to stiffen thermoplastics, for instance polypro-
pylene, polyamides and also more technical polymers such as polybutylene
terephthalate. Thermoplastic processors do not compound themselves their
materials and consequently there is quite a large choice of ready-to-be-
­processed SGF reinforced thermoplastic composites. Table 7.6 gives the aver-
age properties of typical commercial PP–SGF composites, as compiled from
manufacturers’ data sheets (when available).
As can be seen, the hardness, the tensile and flexure properties and the
thermal resistance (HDT, Vicat) generally increase with higher fiber loading,
with however a large scatter that likely reflects the various grades of polypro-
pylene used, as well as (unknown) differences in compounding processes.
E-glass fibers are the most common type but the actual fiber dimensions are
Table 7.6
Commercial PP–SGF Composites; Average Suppliers’ Data
Short Glass Fibers (%wt) 0 10 15 20 30 40 50 60
3
Calculated density (g/cm ) 0.962 0.989 1.015 1.064 1.111 1.155 1.196
Volume fraction from calculated density: 0.0339 0.0500 0.0656 0.0952 0.123 0.1492 0.1738
Volume Fract from given density: 0.0335 0.0581 0.0717 0.1082 0.1406 0.1718 0.2166
Property Unit
Polymers and Short Fibers

Density g/cm3 0.90 ± 0.02 0.95 ± 0.04 1.15 ± 0.212 1.11 ± 0.24 1.21 ± 0.2 1.27 ± 0.2 1.33 1.49 ± 0.04
Melt flow g/10 min 28 ± 18 4.6 ± 3.9 9.3 ± 8.1 11.8 ± 15.0 8.5 ± 11.3 5.1 ± 4.9 – –
Hardness, rockwell R 63 ± 37 84 ± 11 110 95 ± 20 98 ± 17 101 ± 17 – –
Hardness, shore D 60 ± 18 51 ± 26 73 73 ± 3 75 ± 1 79 ± 6 – –
Tensile strength, ultimate MPa 26.6 ± 7.0 34.3 ± 12.7 49.9 ± 8.3 22.9 ± 16.8 69.4 ± 55.6 98.3 ± 52.7 63.0 –
Tensile strength, yield MPa 26.2 ± 8.2 32.0 ± 15.6 60 57.6 ± 29.7 73.1 ± 34.8 87.8 ± 38.4 – 164 ± 34
Elongation at break % 169 ± 83 11.7 ± 11.8 3.3 ± 1.1 4.6 ± 2.8 5.5 ± 7.1 4 ± 4.1 1 1.4 ± 0.5
Elongation at yield % 12 ± 11 3.6 ± 0.8 5 5.7 ± 4.7 4.5 ± 3.1 3.8 ± 2.9 – –
Tensile modulus GPa 1.19 ± 0.35 2.46 ± 0.67 3.48 ± 0.05 4.66 ± 2.32 6.02 ± 3.02 8.06 ± 3.12 11.72 19.00 ± 6.05
Flexural modulus GPa 1.15 ± 0.42 1.87 ± 0.6 3.19 ± 0.61 3.34 ± 2 5.36 ± 3.42 7.1 ± 2.8 8.90 18.07 ± 3.61
Flexural yield strength MPa 30 ± 11 49 ± 15 79 ± 14 71 ± 43 119 ± 81 156 ± 90 98 328 ± 95
Izod impact, unnotched J/cm 9.97 ± 8.56 4.01 ± 1.17 4.54 ± 1.13 3.76 ± 2.65 4.13 ± 3.42 4.61 ± 3.47 1.87
Charpy impact, notched J/cm2 – 0.23 0.23 0.53 1.27 2.04 – –
Gardner impact J 12 ± 11 0.68 ± 0.48 4.51 1.04 ± 0.94 5.00 ± 7.46 0.57 ± 0.52 – –
Izod impact, notched J/cm 2.34 ± 3.83 0.44 ± 0.21 1.31 ± 1.17 1.59 ± 1.96 1.39 ± 1.5 2.75 ± 3.00 0.53 3.78 ± 1.43
Deflection temperature at 0.46 MPa, °C 90 ± 25 117 ± 45 174 ± 40 139 ± 29 150 ± 17 156 ± 9 – –
Deflection temperature at 1.8 MPa, °C 52 ± 6 86 ± 37 129 ± 30 120 ± 42 123 ± 42 142 ± 25 121 318
Vicat softening point °C 108 ± 50 104 ± 12 110 ± 42 132 ± 30 131 ± 35 134 ± 4 – –
359
360 Filled Polymers

20 20
PP + SGF, various suppliers
18 18
PP + SGF , single supplier
16 16
Flexural modulus, GPa

Tensile modulus , GPa


14 14
12 12
10 10
8 8
6 6
4 4
2 2
0 0
0.00 0.05 0.10 0.15 0.20 0.00 0.05 0.10 0.15 0.20
Filler volume fraction Filler volume fraction

Figure 7.11
Mechanical properties of commercial polypropylene–short glass fibers composites; squares
are averaged suppliers data; the vertical bars indicate the standard deviations; shaded dia-
monds are data from one single manufacturer; the curves were calculated with the modified
Guth and Gold equation, using an anisometry factor equal to 4.5.

not disclosed (if known) by the suppliers. The trends observed are therefore
quite remarkable and very significant where tensile and flexural moduli are
concerned (see Figure 7.11). A simple modified Guth and Gold equation with
an anisometry factor equal to 4.5 fits the data well.
Polyamides are particularly versatile polymers, most widely used as
engineering thermoplastic materials. They are melt processable and offer
a unique combination of high mechanical strength, low wear and abrasion
properties together with good chemical resistance. Their semicrystalline
nature results in an excellent combination of properties, which can be fur-
ther reinforced when preparing composites with short glass fibers. Various
types of polyamide are available commercially today including polyamide 6
(polycaprolactame), polyamide 66 (polyhexamethylene diamine adipamide),
polyamide 11 (polyaminoundecanoic acid) and aromatic polyamides, which
can be modified in a variety of ways, through compounding with various
additives including flame retardants, plasticisers, stabilizers, lubricants,
nucleants, mineral fillers, and short glass fibers.
The filling of aliphatic polyamides by short glass fibers is widely used to
improve their physicomechanical and antifrictional properties. Properly
dispersed, glass fibers significantly reinforce the polymer matrix since they
form an internal structure which acts as a load-carrying frame and supports
the main part of the load applied to the sample; therefore higher stiffness,
strength, and heat distortion temperature are obtained, as illustrated by
average properties of typical commercial polyamide–SGF composites. Tables
7.7 through 7.9 were compiled from available manufacturers’ data sheets for
composites based on polyamide 6, 66, and 11, respectively.
Polymers and Short Fibers 361

12
Charpy impact, unnotched, J/cm2 2.4
2.2

Charpy impact, notched, J/cm2


10 2.0
1.8
8 1.6
1.4
6 1.2
1.0
4 0.8
PA6 + SGF 0.6
2 PA66 +SGF 0.4
0.2
0 0.0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Fibers volume fraction Fibers volume fraction

Figure 7.12
Impact resistance of commercial short glass fibers-filled polyamides.

Despite the inevitable scatter due to the diversity of the sources, clear
trends can be seen: mechanical (tensile and flexural) properties, impact resis-
tance as well as heat distortion temperature increase with higher fibers con-
tent. The improvement in impact resistance seems however to be maximum
at around 0.18–0.20 fibers fraction, and damaged samples (i.e., notched) are
considerably more fragile than intact ones Figure 7.12).
Figures 7.12 through 7.14 are plots of flexural and tensile moduli data
for commercial SGF-filled composites based on polyamides 6, 66, and 11,
respectively. It is interesting to compare such data with the predictions of
micromechanical models. With respect to the discussion in Section 7.2, the
Halpin–Tsai model is the most interesting one and the easiest to implement
because all parameters are relatively accessible. Suppliers do not give infor-
mation about the type of glass fibers they used but E-glass type, whose
tensile and flexural moduli are in the 75 GPa range, is the most common
one. The overall fiber orientation in the samples, whose (averaged) prop-
erties are compiled in Tables 7.7 through 7.9, is obviously unknown but
because injection molding is a standard practice in preparing test samples,
near longitudinal orientation is a reasonable hypothesis. This commands to
use Equation 7.4 with ξ = 2(L/D). The fiber aspect ratio is however another
missing information but would the initial L/D ratio be known, compound-
ing and processing operations inevitably break the fibers, in a random and
unknown manner, as documented by some authors.27–29 It follows that the
fiber aspect ratio has to be considered as a fitting parameter whose value
is granted by the best superposition of model curves with experimental
data. Curves in Figures 7.13 through 7.15 were consequently calculated
with Equation 7.4, Efib = 77 GPa, either the tensile or flexural moduli of the
polyamides given in Tables  7.7 through 7.9 and the best fit for L/D. The
upper bound composite moduli were also calculated with Equation 7.1 for
comparison.
362 Filled Polymers

PA 6 + SGF Halpin–Tsai (L/D=50) Upper bound


22 22
20 20
18 18
16 16
Flexural modulus, GPa

Tensile modulus, GPa


14 14
12 12
10 10
8 8
6 6
4 4
2 2
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Fiber volume fraction Fiber volume fraction

Figure 7.13
Flexural and tensile moduli of commercial SGF-filled polyamide 6 composites compared with
the predictions of the Halpai–Tsai model.

PA66 + SGF Halpin–Tsai (L/D=20) Upper bound


20 20
18 18
16 16
Flexural modulus, GPa

Tensile modulus, GPa

14 14
12 12
10 10
8 8
6 6
4 4
2 2
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Fiber volume fraction Fiber volume fraction

Figure 7.14
Flexural and tensile moduli of commercial SGF-filled polyamide 66 composites compared with
the predictions of the Halpai–Tsai model.
Polymers and Short Fibers 363

PA11 + SGF Halpin–Tsai (L/D=20) Upper bound


20 20
18 18
16 16
Flexural modulus, GPa

Tensile modulus, GPa


14 14
12 12
10 10
8 8
6 6
4 4
2 2
0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.00 0.05 0.10 0.15 0.20 0.25
Fiber volume fraction Fiber volume fraction

Figure 7.15
Flexural and tensile moduli of commercial SGF-filled polyamide 11 composites compared with
the predictions of the Halpai–Tsai model.

As can be seen, experimental data are always below, but not far from, the
upper bound curves. This justifies the choice of the Halpin–Tsai equation to
calculate the longitudinal tensile modulus E11 for short fiber-filled compos-
ites. Depending on the polyamide matrix, different values for the (average)
fibers aspect ratio must be used however for the best superposition of calcu-
lated curves on experimental data, i.e., L/D = 50, 20 and 20 for PA6, PA66 and
PA11, respectively. Assuming that fibers had initially the same aspect ratio
before compounding, this would suggest that more breakage occurs in PA
66 and 11 than in PA 6.
Filling polyamides with short glass fibers appears to be a convenient man-
ner to tailor a number of interesting properties for engineering applications.
There are however some substantial disadvantages. For instance SGF-filled
polyamides exhibit quite complex rheological properties, with strong flow-
induced orientation effects and the associated important stress-overshoot
phenomena in processes, such as injection molding.30 Obviously glass-filled
polyamides have a higher abrasivity not only in the molten state which hin-
ders their processing and gives rise to increased wear of processing equip-
ment, but also in the solid state which produces severer wear of adjacent
metal parts in contact with SGF-polyamide joints of friction. Such effects are
common to most SGF-filled polymers but likely exacerbated in the case of
polyamides by the very low viscosity in the molten state. Residual stresses in
molded SGF–polyamide composites are generally higher and whilst a higher
impact resistance is imparted by the fibers (up to a certain level however), a
364

Table 7.7
Commercial PA6-SGF Composites; Average Suppliers’ Data
Short Glass Fibers
(%wt) 0 15 18 20 25 30 35 40 45 50 60
Calculated density 1.11 1.20 1.22 1.23 1.25 1.28 1.30 1.33 1.35 1.37 1.41
(g/cm3)
Volume fraction from 0.0000 0.0606 0.0719 0.0792 0.0971 0.1143 0.1308 0.1468 0.1621 0.177 0.2052
calculated density:
Volume fraction from 0.0000 0.0622 0.0733 0.082 0.1016 0.1208 0.1397 0.1606 0.1816 0.2016 0.2413
given density:
Property Unit
Density g/cm3 1.11 1.23 1.24 1.27 1.31 1.35 1.39 1.45 1.51 1.56 1.66
Tensile MPa 69 ± 21 116 ± 7 140 140 155 ± 7 180 ± 9 192 ± 5 202 ± 6 210 ± 7 226 ± 5 235
strength,
ultimate
Tensile MPa 69 ± 21 116 ± 7 140 140 155 ± 7 180 ± 9 193 ± 5 202 ± 6 210 ± 7 226 ± 5 235
strength,
yield
Elongation at % – 3.3 ± 0.5 4 4.5 3.4 ± 0.2 3.6 ± 0.2 3.1 ± 0.4 3 3 2.5 ± 0.4 2.2
break
Modulus of GPa 2.5 ± 0.6 5.9 ± 0.2 6.90 7.5 8.3 ± 0.3 9.5 ± 0.1 10.9 ± 0.4 12.8 ± 0.3 14.3 ± 0.4 16.0 ± 0.9 20
elasticity
Flexural GPa 2.4 ± 0.6 5.1 ± 0.7 6.80 – 7.4 8.7 ± 0.4 10.2 ± 0.4 10.9 ± 0.3 12.5 13.9 ± 0.6 16.5
modulus
Flexural yield MPa 100 ± 21 163 ± 6 – – 215 235 255 275 330 ± 32 310 340 ± 17
strength
Filled Polymers
Charpy J/cm2 0.85 ± 0.23 0.64 ± 0.08 0.7 1.5 1.15 ± 0.21 1.40 ± 0.10 1.60 ± 0.07 1.77 ± 0.06 1.95 ± 0.07 1.73 ± 0.26 1.5
impact,
notched
Charpy J/cm2 – 3.68 ± 0.79 5.5 – 8 ± 0 9.2 ± 0.4 9.7 ± 0.56 9.67 ± 0.29 10 9.68 ± 0.83 9
impact,
unnotched
Izod impact, kJ/m2 6.8 ± 2 5 8 – – 14 15 – – 17 –
notched
(ISO)
HDT at 0.46 °C – 215 – 215 215 220 220 220 220 220 220
Polymers and Short Fibers

MPa
HDT at 1.8 °C 73 ± 9 191 ± 8 180 195 205 211 ± 6 210 210 210 213 ± 3 210
MPa
Vicat °C 108 ± 50 104 ± 12 110 ± 42 132 ± 30 131 ± 35 134 ± 4 – – – – –
softening
point
365
366

Table 7.8
Commercial PA66-SGF Composites; Average Suppliers’ Data
Short Glass Fibers (%wt) 0 10 13 15 20 25 30 33 35 50
Calculated density (g/cm3) 1.14 1.20 1.22 1.23 1.26 1.28 1.31 1.32 1.33 1.40
Volume fraction from calculated
density: 0.0423 0.0543 0.0622 0.0812 0.0995 0.1171 0.1272 0.1340 0.1810
Volume fraction from given
density: 0.0426 0.0548 0.0627 0.0833 0.1023 0.1216 0.1337 0.1417 0.2028
Property Unit
Density g/cm3 1.14 1.21 1.23 1.24 1.29 1.32 1.36 1.39 1.41 1.57
Tensile strength, ultimate MPa – 105 – 125 145 170 190 130 210 240
Tensile strength, yield MPa 90 105 100 125 – – – – – –
Elongation at break % – 4 – 4 4 3 3 4 3 2
Elongation at yield % – – 4 – – – – – – –
Modulus of elasticity GPa 3.1 5.0 5.5 6.2 7.4 8.4 9.7 8.5 11.7 16.2
Flexural modulus GPa 3.0 4.5 – 5.3 6.4 7.3 9.1 7.8 9.5 13.5
Flexural yield strength MPa 125 – – – – 280 280 – – –
Charpy impact, notched J/cm2 0.4 0.5 0.6 0.7 0.8 0.9 1 1.4 1.4 1.6
Charpy impact, unnotched J/cm2 3.5 6.5 4.1 4.8 5.4 8 8 9.5 9.5
Izod impact, notched (ISO) kJ/m2 4.5 5 8 6 7 8.3 10 15 13 14.5
HDT at 1.8 MPa °C 85 232 228 245 250 255 253 248 255 255
Filled Polymers
Table 7.9
Commercial PA11–SGF Composites; Suppliers’ Data
Short Glass Fibers (%wt) 0 10 15 20 23 30 40 43
Polymers and Short Fibers

Calculated density 1.06 1.12 1.15 1.17 1.19 1.23 1.27 1.29
Volume fraction from calculated density 0.0000 0.0395 0.0580 0.0759 0.0863 0.1097 0.1412 0.1501
Volume fraction from given density 0.0000 0.0391 0.0571 0.0969 0.0884 0.1127 0.1495 0.1655
Property Unit
Density g/cm3 1.06 1.11 1.13 1.50 1.22 1.26 1.35 1.42
Tensile strength, ultimate MPa 38.0 82.7 68.9 110.0 110.0 89.6 82.7 146.0
Elongation at break % 50.0 4.0 4.5 5.0 5.0 5.3 3.5 4.0
Tensile modulus GPa 1.30 2.76 3.79 6.69 7.69 7.45 8.27 12.40
Flexural modulus GPa 1.90 2.76 3.03 5.52 – 5.52 6.89 –
Flexural yield strength MPa – 96.5 96.5 117.0 – 131.0 124.0 –
Izod impact, unnotched J/cm 5.87 5.34 6.41 – 7.47 4.54
Izod impact, notched J/cm – 0.747 0.534 1.070 – – 1.120 –
HDT, at 0.46 MPa °C 135 163 178 179 – – 188 –
HDT, at 1.8 MPa °C 47 154 170 174 – 182 182 –
367
368 Filled Polymers

lower toughness is observed in impacted regions due to damages arising


around the fibers in the matrix . These disadvantages can be somewhat com-
pensated by the introduction of various modifying ingredients into glass-
filled polyamides.

7.4 Typical Rheological Aspect of Short Fiber-filled


Thermoplastic Melts
As we have seen, short glass fiber-thermoplastic polymers offer improved
stiffness, strength, and heat distortion temperature with respect to the
unfilled polymer. But the overall properties of the fabricated parts strongly
depend on the interrelation between macroscopically observable rheologi-
cal properties, the local fiber orientation in the processing fields, and the
resulting mechanical and thermal properties in the solid state. As clearly
demonstrated by the mechanical models discussed above, the optimum
reinforcement is only obtained when the fibers are properly oriented. How
fibers orient themselves during flow is a key factor, obviously depending of
the matrix flow properties, its shear and extensional viscosities and its shear
thinning character, and the complex flow fields of the final processing opera-
tions (i.e., before the cooling down and the associated recrystallization step,
if any), notwithstanding of course the elasticity of the fluid which surely has
an influence on the orientation of the fibers.
In contrast with the industrial importance of short fiber-filled thermoplas-
tics, there is a rather limited scientific literature on the rheological properties
and behavior of such systems and only a few key rheological experiments
were performed with model systems. Amongst the likely reasons, there are
the tremendous experimental difficulties in correctly assessing the flow
properties of molten short fiber–polymer composites, due to a number of
singular effects, e.g., nonlinear Bagley plots in capillary rheometry,31 nonre-
peatability of steady shear experiments with cone-and-plate, rheometers, 32
orientation of the fibers during rheometrical test, nonlinear effects, 33 etc.,
Around 30 years ago, Chan et al. reported data on short glass fiber filled
polyethylene and polystyrene melts34,35 and Knutson et al.36 investigated
the extrusion behavior of SGF-filled polycarbonate. The orientation of the
fibers under various flow conditions was investigated by various authors,
namely Laun,30 Crowson and Folkes,31,37 and S. Kenig,38 who proposed a
qualitative model as well as an identification of the various flow mecha-
nisms that likely contribute to fiber orientation during injection molding,
i.e., spreading radial flow in the vicinity of the cavity gate, fountain flow in
the advancing front, converging type flow in the melt front zone, and shear
flow at some distance from the wall. As a result of these mechanisms a lay-
ered structure is formed, having distinct fiber orientations. Other authors
Polymers and Short Fibers 369

reached similar conclusions, for instance Vincent and Agassant39 with the
injection molding of a SGF–polyamide composite in center gated discs and
Larsen40 with the injection molding of a rectangular box using short fiber
reinforced polypropylene.
In steady slit extrusion of short fiber reinforced thermoplastic composites,
the formation of a skin-core of fibers oriented in the thickness direction has
been observed, with fibers axis along the flow direction and parallel to the
walls in the skin region, irrespective of the entrance geometry. Different fiber
orientation distributions in the core region are however obtained by using
different entrance geometries.41
During processing, fibers move, tumble, and rotate with the flow of the
polymer matrix, which inevitably changes their orientation. Providing, the
loading is sufficiently low for no or minimum interactions to take place
between neighboring fibers, it is easy to understand the basic difference in
short fibers motion in either extensional or shear flows. In (simple) shear
flow, an isolated fiber tumbles over and over as it follows the mainstream, as
foreseen by Jeffrey.42 In extensional flow, the same fiber rotates until its long
axis is aligned with the main extensional axis and then stays there.43–45 In
addition, a fiber in a shear flow has a minimum effect on the flow, when its
axis is approximately parallel to the streamlines, and therefore, whilst tum-
bling, tends to stay in this position. In contrast, a fiber in extensional flow
field is permanently in the position for which it has the maximum effect on
the flow. Such behavioral differences explain why dilute suspensions of rod
like particles and fibers exhibit substantial deviations from the Trouton ratio
of 3. It must be noted that most published studies have been performed on
model systems, with for instance short fibers in either Newtonian fluids or
viscoelastic polymer solutions. One presumes that the behavior of the fibers
in molten polymers is similar.
Generally, converging flow results in high fiber alignment along the flow
direction, whereas diverging flow causes the fibers to align at 90° to the
major flow direction. Simple shear flow produces a decrease in alignment
parallel to the flow direction and the effect is more pronounced at low flow
rates. The net overall fiber orientation in a given process therefore results
from the particular combination of converging-diverging and shear flow
fields prevailing in the final shaping steps, i.e., the die in extrusion or the gate
and the cavity in injection molding, before the melt solidifies. In addition to
flow induced orientation effects, that are particularly strong in extensional
flow fields, another but less understood mechanism is associated with fiber-
to-fiber interaction. Indeed, in many practical applications, the fiber volume
fraction is sufficiently high for fibers to flow in close proximity to each other.
In such a case, various types of fiber interactions may come into play, for
instance excluded-volume (due to fiber tumbling and/or rotation), and fric-
tion or mechanical interactions between fibers, in addition to usual hydro-
dynamic forces effects. A recent work by Guo et al.29 on E-glass fibers-filled
(up to 35%) linear low-density polyethylene melts generated experimental
370 Filled Polymers

results that clearly demonstrate the effects of fiber–fiber interactions and the
coupling between fiber orientation and polymer chains conformation on the
rheological properties of molten composites.

7.5  Thermoplastics and Short Fibers of Natural Origin


Certain natural fibers such as jute, hemp, flax, sisal, etc., have relatively high
strengths that make them attractive for preparing composites for various
applications. Like any other fibers, aspect ratio, loading, and orientation are
the first parameters that control the properties of the natural fibers com-
posites and, providing differences in mechanical properties (e.g., moduli)
between the fibers and the polymer matrix are sufficiently large, similar
reinforcing effects can be expected as with their synthetic counterparts.
However the efficiency and the level of reinforcement may be different,
due to a few specific aspects. Firstly, natural fibers often exhibit quite large
degrees of nonuniformity in most characteristics, i.e., chemical composition,
surface aspect and properties, length and diameter, cross-sectional shape,
strength, and stiffness.46 In addition, most cellulose based materials are
prone to moisture absorption and therefore the durability of the composite
may be questionable. Most polymers and hence polymer composites absorb
moisture in humid atmosphere or when immersed in water. Natural fibers
absorb of course more water than synthetic fibers, which results in degrada-
tion of mechanical properties such as tensile strength. Moisture diffusion
in a composite depends on various factors, with fibers volume fraction, void
volume, fibers treatment and additives, humidity and temperature the most
important ones.
Preparing polymer composites with lignocellulosic fibers is not just a nov-
elty. Contemporary environmental concerns and the growing concern for
sustainable development have renewed the interest in materials that offer
both economic and environmental benefits. Applications considered are in
the automotive, construction, furniture and packaging industries. Chopped
natural fibers, with controlled aspect ratios and tailored specific proper-
ties, could be used for higher value applications than typical wood polymer
composites. An attractive aspect is that compared to short glass fibers, natu-
ral fibers are nonabrasive and consequently higher fiber content in plastics
could be considered, without excessive wear and damage of compounding
and processing equipment. The high moisture sensitivity of natural fibers
and their low microbial resistance are however disadvantages that needs to
be considered, namely by insuring that they are properly encapsulated in
the polymer with good fiber-matrix bonding, either using the appropriate
additives or through suitable preliminary treatments, for instance through
acetylation of hydroxyl groups present in the fiber.47
Polymers and Short Fibers 371

Like wood flour, the main limitation of the use of natural fibers is the lower
permissible processing temperature, in order to avoid fiber degradation and
/ or emission of volatiles. The processing temperatures are thus limited to
180–200°C, although it is possible to use higher temperatures for very short
periods. Of course, processing equipment must be properly designed in
order to avoid any possibility of stagnation or low flow zones. Practically,
this limits the type of polymers that can be used to commodity thermoplas-
tics such as PE, PP, PVC and PS.
The main component of natural fibers is cellulose that, owing to its crystal-
line structure, has one of the highest tensile modulus exhibited by polymer
materials, around 136 GPa compared with 77 GPa for E-glass fibers. Inside a
natural fiber, cellulose chains are assembled in microfibrils which are packed
in several layers. The high stiffness of natural fibers is due to the helicoidal
supramolecular organization of crystalline cellulose which, during stretch-
ing however brings a torsion/traction coupling that can modify the fiber–
matrix interface and therefore affects the deformation behavior and the
ultimate fracture mechanisms. There are also hydrogen bonding between
cellulose chains which explain the excellent properties of cellulose based
fibers.48,49 All these characteristics explain the large diversity in mechanical
and physical properties of natural fibers.
Table 7.10 somewhat positions typical natural fibers such as jute and kenaf
vs. glass fiber, when used at equal level in polypropylene. Tensile, flexural,
and impact properties, along with water absorption and specific gravities
were measured on injection molded samples. As can be seen, natural fibers
composites are quite challenging the glass fibers-filled composite with how-
ever a significant difference in moisture sensitivity.
So-called WPCs are thermoplastic composites that combine wood flour
with plastics, intended to offer the advantages of wood-like materials with
the processing and part-design capabilities of thermoplastic polymers.

Table 7.10
Commercial Fibers-Filled Polypropylene Composites
PP-fibers Virgin PP E-Glass Jute Kenaf
Fibers content %wt 0 50 50 50
Fibers volume fraction 0 0.15 0.39 0.39
Specific gravity g/cm3 0.90 1.33 1.08 1.07
Tensile modulus GPa 1.7 11.5 8.1 7.9
Tensile strength MPa 30 88 69 58
Elongation at break % 170 1.4 2.3 2.2
Flexural modulus GPa 1.2 8.9 7.3 7.3
Flexural strength MPa 40 98 97 98
Water absorption %/24 h 0.02 0.05 1.13 1.05
Linear mold shrinkage cm/cm 0.028 0.004 0.003 0.003
372 Filled Polymers

Whilst finely grinding wood material may considerably hide the fiber aspect
of wood particles, WPC exhibit properties which are typical of systems filled
with short fibers of natural origin. As WPC is a rich subject in itself, only a
few basic aspects will be discussed hereafter. It is obvious that not all thermo-
plastics can be used to prepare WPC. Indeed the polymer component melts
or at least softens at or below the degradation point of the wood component,
normally 200–220°C but must be rigid at normal usage temperatures (i.e., up
to around 65°C). Practically, this reduces the choice to polypropylene, low-
and high-density polyethylenes, polystyrene, and vinylics (essentially PVC).
As such wood flour, used as a filler in thermoplastic composites, offers
only modest, if any, reinforcement, but wood fibers can lead to superior
composite properties and act more as reinforcing filler. Commercial wood
flour is a by-product of the wood industry, often mechanically processed
from waste materials such as planer shavings, chips, and sawdust, which
are reduced to fine powders, with various grades available depending upon
the particle size and the wood species. Wood (cellulose) fibers are produced
through more or less complex defibrillation techniques, using raw materi-
als from both virgin and recycled resources, and are different from natural
fibers, such as jute, hemp, or sisal.
Composites with wood material and polymer can be prepared in two ways.
In the first, the wood fiber/flour is a reinforcing agent or a filler in a continu-
ous thermoplastic matrix. In the second, the thermoplastic is a binder to the
majority wood component. Only the first approach is discussed hereafter
since a continuous thermoplastic matrix determines the processability of the
composite material and is the necessary condition for processing WPC with
conventional thermoplastic processing techniques, whilst certain adaptation
of the equipment and operation may be required.
Wood-filled thermoplastic composites were introduced on the market in
the early 1990s, essentially in North America. By far the biggest sector of
the WPC industry is the fabrication of deck board and railing, fencing, door
and window frames for the residential construction market. Deck board,
alone, accounts for about 60% of overall WPC volume, with a steady growth
of nearly 14%/year from 1996 to 2006, despite the 10–20% higher price com-
pared to treated wood. Obviously in the long run, WPC need less mainte-
nance than natural wood and the initial higher investment is recovered in
a few years. The second sector is the fabrication of wood-filled thermoplas-
tic composite lumbers, with a similar growth rate. It is worth noting that
all such products are made through continuous process, with extrusion the
major technique. Natural orientation of wood fibers along the main flow lines
in suitably designed extrusion dies is likely the reason as it will optimize
the mechanical performance (in terms of flexural and tensile modulus, for
instance) and aesthetics (surface aspect). Injection molding and compression
molding together account for less than 10% of the WPC market, with more
developments in Europe (Austria and Germany) than in North America.
Injection molding is a challenging technique for WPC because wood begins
Polymers and Short Fibers 373

to degrade at temperatures above 200°C and flow easiness dramatically


decreases with increasing wood flour/fiber content. WPC cannot be consid-
ered only as mere substitutes to natural wood because using the resources of
thermoplastic processing techniques, products can be fabricated that would
be impossible or at least difficult and costly to make with wood. Indeed
extrusion allows long, hollow profiles with any cross-sectional design to be
made. The resources of coextrusion and other specific techniques can also
be exploited, for instance by introducing a foaming agent into the molten
wood/plastic matrix, so that lighter products with an internal rigid foamy
structure are obtained.
A basic problem in the production and processing of wood polymer com-
posites is the moisture content and sensitivity of natural materials, which
generally require a thorough drying before compounding and a control of
the storage conditions of the compounds before the ultimate shaping/pro-
cessing steps. It is generally considered that the maximum moisture content
for easy processing is around 10–12% for extrusion or calendering opera-
tions, and less than 3–4% for injection molding. Generally speaking, a thumb
rule is the drier, the better.
When developing WPC, a number of technical problems must be solved,
for instance fiber-to-matrix bonding, fiber damage during compounding and
processing, fiber orientation in finished part, providing obviously that the
correct fiber-matrix selection has been made. Basic research works to tackle
one or several of these issues have been initiated around 40 years ago, with
varied industrial or scientific successes, and still go on owing to the great
number of likely interesting systems and the recently renewed interest in
materials of natural origin. Ageing is an issue in WPC technology and vari-
ous additives are used to protect the wood flour or fibers against moisture,
light, mold, mildew, insects, and other pests. But the best technique is a full
encapsulation of the wood particles or fibers in the polymer matrix.
Virgin and recycled PE (both high- and low-density) remains the most
common thermoplastic used in the WPC industry. PP offers enhanced prop-
erties (higher stiffness, less creep, higher HDT) but its higher melt temper-
ature may pose some processing challenges. The use of PVC continues to
grow as more structurally enhanced products emerge. Although most WPCs
still incorporate wood flour, derived from reclaimed pine, maple and oak
sawdust, planer shavings, sanding dust, and milled scraps, there are today
alternatives that permit more targeted products in terms of performance,
for instance pulp cellulose fiber, which is generally reclaimed from post-
consumer waste paper.50 Pulp cellulose fiber has zero coefficient of thermal
expansion and, because it contains no lignin, it does not photo-degrade like
wood flour. Natural fibers are also an alternative but large scale uses require
a regularity in large volume availability, at constant quality. The potential
for natural fiber is likely as a reinforcing component with wood floor the
major filler. Flax, jute, hemp, and kenaf fibers can be produced with millime-
ter range lengths and, properly oriented in the material would give enhanced
374 Filled Polymers

flexural properties. Compared to short glass fibers, natural fibers have higher
strength-to-weight ratios, similar surface functionality and handling charac-
teristics as wood, and can be easily incorporated into a wood composites.51
Today a wood–plastic composite is quite a complex formulation. A typical
WPC formulation would consist for instance of 30–35%wt polymer, 50–55%
wood flour, 5% talc, 4% lubricant, 2% colorant, 2–3% other additives such as
coupling agents and/or mold and tannin inhibitors. Additives play a key role
in enhancing product characteristics, particularly with respect to weather-
ability. Additives are lubricants, UV stabilizers, flame retardants, antimicro-
bials, color concentrates, coupling agents, foaming agents, dispersion agents,
mineral fillers, antioxidants, and compatibilizers. Maleic anhydride-based
coupling agents, for instance MA-grafted polyolefins, improve dispersion
and wetting of the lignocellulosic materials by the polymer, with a direct
effect on product durability, because the rate of moisture absorption is
reduced, while strength properties are improved.52 Lubricants improve sur-
face appearance and processability, and reduced manufacturing cost through
increased throughput. Zinc borate is quite a common additive to WPCs to
ward off fungal decay and inhibit the growth of mold and mildew.
A number of published research works allow to understand some basic
scientific aspects of the WPC technology.53–55 The effect of wood flour par-
ticle size is qualitatively the same as with other particulate fillers, except
that the fiber nature seems to play in role in what the ultimate properties
are concerned, as illustrated in Figure 7.16 with results on (pine) wood flour-
filled polypropylene composites.56 As can be seen, tensile and flexural mod-
uli increase with increasing wood flour content (left graph); whilst flexural
strength is practically not affected, the tensile strength steadily decreases
with increasing filler content (right graph). In the range considered, 24–282
µm, particle size has practically no effect.

Polypropylene/wood flour
Tensile or flexural strength, MPa

5 50
Tensile or flexural modulus, GPa

Flexural
Tensile
4 40
Flexural Tensile
3 30

2 20
Part.size, µm
1 282
10
24
0 0
0 10 20 30 40 50 0 15 25 40
Wood flour content, % Wood flour content, %

Figure 7.16
Effect of (pine) wood flour on the mechanical properties of PP based composites. (Data from
J.C. Caraschi, A.Lopes Leão. Mat. Res., São Carlos, Brazil, Oct./Dec.5 (4), 405–409, 2002.)
Polymers and Short Fibers 375

There are generally large differences in surface polarity between wood


particles or fibers and polyolefin matrices, which lead to poor wetting, hence
dispersion mixing difficulties and eventually limited mechanically proper-
ties. Various coupling agents have been studied in the literature with some
advantages assigned to functionalized polymers.57 Maleic anhydride-grafted
PP, PE-PP and (in a lower extent) PE are the most commonly used, likely
owing to their wide commercial availability, e.g., Exxelor® VA1801, MA-g-
EPR, 1.21%MA (Exxon Mobil Chemical Co) or Epolene® G2608, MA-g-PE,
acid number: 8 mgKOH/g (Eastman Chemical Co). Such compatibilizers
readily enhance the mechanical properties of composites, especially tensile
and flexural strength, likely because on one hand either polar interaction or
covalent links are obtained between the anhydride carbonyl and hydroxyl
groups of wood surfaces, and on the other hand because of good compatibil-
ity with polyolefins.58,59 Stark and Rowlands60 studied for instance the effects
of 40%wt wood flour and fiber of different particles size (aspect ratio), com-
pounded in polypropylene, without or with maleated PP as coupling agent.
Tensile and flexural strength and modulus were found to increase with the
fiber aspect ratio. Notched impact energy increases with higher particle size,
whereas unnotched impact energy exhibits the reverse trend. Wood fibers
give generally higher strengths than wood flour, but the higher aspect ratio
had little effect on impact energy. The MA-grafted PP coupling agent gives
higher strength and the effect is larger with wood fibers than with wood
flour. The coupling agent does not significantly affect tensile or flexural mod-
uli. Dányádi et al.61,62 studied the effect of MA-grafted PP ­molecular weight
and content on the properties of wood flour filled polypropylene composites.
They found that stiffness increases with wood flour content, with very little
effect from the type or the amount of the functionalized polymer additive
used. However, ultimate properties are strongly influenced by the amount
and properties of MA-g-PP, with larger molecular weight and smaller func-
tionality having beneficial effects on both tensile strength and impact resist-
ance. Indeed, the functionalized agent increases considerably the interfacial
adhesion between the wood particles and the polymer matrix, which com-
pletely changes the deformation mechanism; the fracture of wood parti-
cles becomes the dominating factor because the matrix polymer essentially
deforms by shear yielding. The optimum MA-g-PP/wood flour ratio is
around 0.05.

7.6  Elastomers and Short Fibers


Filling an elastomer with short fibers essentially involves combining the vis-
coelastic behavior of a rubber matrix with the strength and stiffness of the
fiber, in order to obtain useful engineering materials. Reinforcing elastomers
376 Filled Polymers

with long fibers, textiles, or fabric is a long established practice in the rubber
industry, namely in tires, belts, hoses, and other demanding applications, but
long fiber rubber composites are outside the scope of this book. It is in prin-
ciple possible to prepare short fiber-filled rubber composites by exploiting
the resources of both the rubber latex and the regenerated cellulose technolo-
gies. Research results have been published that described such materials.63–65
For instance, short-fiber filled compositions were prepared by coprecipitat-
ing mixtures of natural rubber or nitrile rubber latex with cellulose xanthate.
Compounds with cellulose fiber content up to 30 phr were obtained that
exhibit increased tensile modulus and strength and decreased elongation at
break as fiber level increases, as well as some (minor) effects on curing prop-
erties. It is clear that such systems are worth consideration with respect to
their potential for sustainable development, because they partly use renew-
able components. However, despite encouraging laboratory results, there has
been so far no industrial application (to the author’s knowledge). Only more
practical rubber compounds with short fibers added during mixing opera-
tions are discussed hereafter.
Because the dispersed fibers do flow with the rubber matrix during the
processing operations, quite intricate parts can be shaped through tech-
niques such as extrusion, calendering, and the various molding techniques
(compression, transfer and injection), then vulcanized. There are however a
few constraints for fiber-reinforcement to be used in rubber technology: first
an adequate dispersion of fibers must be achieved, with limited fiber break-
age, second a good adhesion between fibers and rubber must be obtained,
generally by using either a suitable pretreatment of the fibers or an appro-
priate adhesion-promoting bonding system/agent. In contrast with fibers-
filled thermoplastics for which a good wetting of the fibers by the polymer
is generally sufficient because only (small strain) elastic properties are con-
cerned, most rubber parts must support larger strain, in the several 100%
range. Rubber-fiber bonding is therefore essential, and can only be achieved
by chemical means.
Since the early works in the late 1970s by Coran et al.,66–68 short natural
fiber–rubber composites have received a considerable attention in scien-
tific literature and a few systems have achieved a significant importance
in certain engineering and consumer goods applications, because of their
high strength-to-weight ratio, manufacturing flexibility, and ease of pro-
cessing. Particularly, these authors demonstrated that in sulfur-cured SBR
compounds, unregenerated (hardwood) cellulose fibers, properly bonded
to the matrix (by an undisclosed system), gives more that 60% of the ten-
sile modulus that would be obtained with the same volume fraction of short
glass fibers. They also concluded that, with respect to the benefits offered
by cellulose fibers, not much advantages would be obtained by using fibers
with modulus higher than around 25–40 GPa. Such results prompted the
development of unregenerated cellulose grades especially for use in elas-
tomer reinforcement,69 which were commercialized under the trade name
Polymers and Short Fibers 377

Santoweb® by Monsanto Co. (now Flexsys). Today, four grades of treated


cellulose fiber product are commercially available for use in NR, SBR, BR,
and CR compounds (Santoweb® D), or EPDM and IIR elastomers (Santoweb®
H). It seems that the treatment consists in adding the correct amount of
resorcinol-formaldehyde resin, i.e., a methylene receptor, so that the use of
a methylene donor chemical, for instance hexa(methoxymethyl)melamine,
HMMM, is required. Maybe the most significant industrial achievements in
rubber technology with such treated cellulose fibers are based on the fact
that, during processing, fibers tend to become oriented in the direction of the
flow. Special designed extruder dies (so-called expanding mandrel extrusion
dies) can therefore be used to provide control of the average fiber orienta-
tion, such that hoses can be made with fibers oriented in the circumferential
direction.
Extensive reviews on short fibers-reinforced elastomers have been pub-
lished by Goettler and Shen,70 and more recently by Rajeev.71 Table 7.11 is a list
of selected published works on natural fiber-filled rubber composites, sorted
by rubber type. Only vulcanizable rubbers were considered in preparing
the table; thermoplastic rubber and rubber-plastics blends were omitted. As
can be seen most grades of conventional elastomers have been considered
with quite a large variety of natural fibers. Natural rubber, SBR, and EPDM
compounds have received much attention, as expected with respect to their
industrial importance. A bonding system is used in most cases, with resin
types (i.e. resorcinol-formaldehyde, resorcinol-hexamethylenetetramine, sil-
ica-resorcinol-hexamethylenetetramine) the most frequent ones.
The predominance of resin bonding is likely related to the thorough
understanding of the chemistry involved when using resin formers to
obtained fiber-to-rubber adhesion. The development of adhesion essentially
occurs In situ, during vulcanization, when resin forms in the boundary
fiber–matrix regions. As demonstrated by Morita,72 typical Mooney scorch
curves are readily observed when resin formation and rubber vulcanization
are made to occur subsequently through adequate choice in the formula-
tion. Recommended formulations are however such that both events occur
simultaneously during curing. Resorcinol or resorcinol derivatives are com-
mon resin formers, playing the role of “methylene acceptors,” while hexam-
ethylenetetramine (HEXA) or HMM act as “methylene donors.” Silica has
been found to enhance the adhesion significantly. As mentioned above, such
resins systems were the background of patents obtained by Monsanto Co
more than 30 years ago. It is pretty obvious that a pretreatment of fibers with
resorcinol or resorcinol-formaldehyde will favor fiber dispersion because it
will reduce hydrogen bonding between fibers and therefore give an easier
separation of the single fibers from fiber bundles, with a minimum force
in such a manner that fiber breakage is limited during mixing. Industrial
usages over the last 20 years have largely demonstrated that resin systems are
likely the best choice for cellulose-lignin fibers, i.e., most natural fibers. Resin
systems are not very effective with polyester fiber, not effective at all with
378 Filled Polymers

Table 7.11
Selected Published Works on Natural Fiber Filled Rubber Composites
(Vulcanizable Rubbers only)
Rubber(s) Fiber(s) Bonding System Reference
CR Cotton, polyester, None 88
Cellulose
CR PET, PA R-F* 89
CR Silk S-R-H* 75
CR, EPDM, PUR PET None 90
EPDM PET 1,4-carboxy- 91
sulphonyl-diazide
EPDM Melamine S-R-H 92,81
EPDM PAN, Aramid None 93
EVA PAN, Carbon None 94
NBR Jute S-R-H 95
NR Jute S-R-H 96, 87, 97
NR Silk S-R-H 98
NR Polyester None 99
NR PA Undisclosed/none 100, 101, 105
NR Grass Silane (TESPT)* 102
NR Hemp None 103
NR, NBR Glass, PA, carbon, S-R-H 74
aramid, cellulose
NR, SBR, EPDM Glass, aramid, None 104
cellulose
NR + LDPE + Liquid NR Kenaf Silane, PPgMA 105
PU, NR latex PA, aramid, Corona, γ-irradiation 106
HMW-PE
SBR Jute, glass R-H* 107
SBR Glass, PA, carbon, R-F 108
cellulose, polyester
SBR Sisal R-H 109
SBR, BR Cellulose R-H 84
XNBR Jute S-R-H 88
* R-F: resorcinol–formaldehyde; R-H resorcinol-hexamethylenetetramine; S-R-H: silica-­
resorcinol-hexamethylenetetramine; TESPT: bis(3-triethoxysilylpropyl)tetrasulfane.

glass fiber and mixed results were reported with silk, amide and aramide
fibers in various elastomer systems.73–75 Special fiber–elastomer composites
have been reported to exhibit good properties without the need for bond-
ing systems, e.g., polyurethane and aramid short fibers.76 It is clear that the
possibility to fine-tune a short-fiber filled rubber compounds through (pro-
prietary) adjustments of the formulation suits well the practice of the rubber
industry. Wennekes et al.77 have recently published quite an extended review
of the various treatment/additives that can be used to promote fiber-rubber
Polymers and Short Fibers 379

adhesion. They described namely other alternatives to the well established


resorcinol-HEXA system, such as fiber surface roughening (through aggres-
sive chemical treatment), adhesion promoter additives, impregnated fibers,
and plasma treatment.
Whatever are the natural fiber and the rubber type, qualitatively the same
results are always reported when increasing the short fiber content:

• Reinforcement of the tensile properties, i.e., increased modulus and


tensile strength, but lower elongation at break
• Increased hardness and stiffness
• Improvements in resistance to cuts, tears and punctures
• Enhanced resistance to solvent swelling (constraining effect of
fibers)
• Lower extrudate swell (essentially reflecting fiber alignment in the
flow direction)
• Overall benefits increase when an appropriate bonding system is
used.

Expectedly quantitative results vary considerably from one fiber–rubber


system to another, and other compounding ingredients may induce addi-
tional effects (positive or negative). The qualitative effects are completely in
line with the expected role of short fibers, in agreement with micromechanic
considerations (see Section 7.2). Certain authors have used well established
micromechanic approaches, e.g., Voigt and Reuss averages (Equations 7.1
and 7.2), and Halpin–Tsai equations (Equation 7.5) to consider the effects on
short natural fibers in rubber compounds.78–80
The trade literature shows that certain short natural fiber–rubber compos-
ites have achieved significant importance in a number of applications. This
is certainly true for fiber materials that are commercially available and that
meet the requirements in product quality and consistency for engineering
applications, e.g., the Santoweb® treated cellulose fibers grades from Flexsys,
the cotton flocks W200 and D220 supplied by IFC, International Fiber Corp.,
North Tonawanda, NY, the “jet process” milled cellulose fiber Interfibe®
­supplied by Interfibe LP, Solon, OH and the nitrile rubber coated cellulose
fiber Nicote® supplied by Vellumoid Inc., Worchester, MA. All those com-
mercially available products are fibers of 10–25 µm diameter and average
length in the 250–600 µm range.
Short fiber reinforced rubber composites have found well established usages
in hose, belt, tires and other automotives applications, generally because
either they offer specific properties that cannot be obtained with more con-
ventional fillers only (e.g., carbon black or silica) or they are a cheaper alter-
native at equivalent properties. It is obviously the capability of short fibers to
be properly oriented during the (adapted) fabrication process and therefore
to achieve well-controlled anisotropic mechanical properties, that is the most
380 Filled Polymers

critical reason for using such systems. Providing the fiber aspect ratio is larger
than 100, an adequate mechanical anisotropy can be imparted to the rubber
composite through the processing operation. Indeed, with high aspect ratio
treated cellulose fibers, large difference in tensile properties are achieved
between the processing flow direction and the cross-machine direction in
carefully milled sheets (up to 10 times in term of tensile modulus).55 Short
fibers, well dispersed in a viscoelastic matrix at concentrations above 6%,
can therefore be very efficiently oriented in certain type of flow fields where
the elongational component is dominant. Goettler and Lambright developed
a proper design of a converging–diverging extruder die such that average
fiber deviation from the flow axis is substantially different.81,82 As illustrated
in the upper part of Figure 7.17, the right combination of (wall) shear forces
with the extensional flow occurring between the screw head and the orifice
of a rubber hose die cause the fibers to become aligned parallel to the flow
direction. Because, at some intermediate point, the restriction is followed by
a specific type of expansion, a somewhat controllable transverse orientation
pattern of the fibers is obtained.
As illustrated in the lower part of the figure, this technique led to the
so-called “extrusion moving die” technology to produce curved hoses.83
By moving the inner or the outer portions of the die out of concentricity
in a programmed sequence, the opening between inner and outer dies
is varied and the extrusion direction can be made to deviate from the
machine axis. S-bend hoses can thus be produced, whose performances
(i.e., burst pressure) are similar to cord reinforced coolant hoses. Die com-
ponent motions and timing sequences must be controlled in a very tight
manner through computer control. This special extrusion technique was
commercially available by the mid 1980s (namely from the now defunct
Iddon Brothers Ltd Co) and is in widespread use today. This unique pro-
cess for producing bend hoses is however contingent upon the use of
short fiber reinforcement, since well dispersed fibers must be present in
the compound before it is extruded. It suits obviously well treated cel-
lulose fibers like Santoweb®. As reported by Goettler,84 the technology is
not restricted to rubber materials but can also be applied with soft ther-
moplastics, such as plasticized PVC. In such a case an isocyanate bonding
agent is recommended.85
Of all the fibers that have (and are still) evaluated in rubber compounds
over the last 30 years, only treated unregenerated cellulose fibers and
chopped aramid fibers have really achieved a certain industrial importance.
The current research efforts being made, essentially by university groups
in the producing countries, on natural fibers such as jute, sisal, kenaf, etc.,
are not (yet) leading to documented industrial applications, at least to the
author’s knowledge. Actual reasons for this situation are unclear, except
maybe that industrial applications for treated unregenerated cellulose fibers
were inherited from the important preliminary works by former Monsanto’s
scientists, prior to the commercial development of Santoweb® fibers. When
Polymers and Short Fibers 381

Converging mandrel die Converging–expanding die

Circumferential Mixed fiber


fiber orientation orientation

Moving extrusion die technology

Mandrel axis

Outer die axis

Lateral translation of outer die produces eccentricity then hose curvature

Figure 7.17
Controlling short fiber orientation in extruded rubber hoses with converging–expanding dies;
application in the moving die technology to produce curve hoses.

comparing the effects of various type of fibers, i.e., hardwood cellulose,


acetate, nylon 66, PET, acrylic and glass, on the mechanical properties on
a carbon black filled SBR compound, Coran et al.54 came to the conclusion
that properly bonded and treated cellulose fibers gave the best balance of
performances, namely tensile properties not so far to what is obtained with
short glass fibers. In fact, when normalizing their results for changes in
fiber concentration, orientation, geometry and matrix modulus, they clearly
demonstrated that not much is gained with fibers whose tensile modulus
is higher than about 27–40 GPa. Glass fibers exhibit typically tensile modu-
lus in the 70–90 GPa range but suffer from severe breakage during mixing
382 Filled Polymers

operations because (like carbon fibers) they have low ­bending strength.
Ductile fibers, such as cellulose and synthetic fibers, are more flexible and
resistant to bending, the former because they are generally ribbon shaped.
Therefore such fibers have the capability to preserve aspect ratios largely
above the critical value for an effective stress transfer from the matrix to
the fibers.
A number of factors must be taken into consideration when comparing
the potential of different fibers for (sizeable) industrial applications. Regular
large volume availability with minimum variation in properties is obviously
required and is quite a challenging constraint for all materials of natural ori-
gin. In addition, there are some important technical constraints to consider,
i.e., to name a few:

• Fiber size and aspect ratio (in the composite, after the mixing
operations)
• Fiber–polymer matrix interactions, with respect to the fiber pretreat-
ment and/or the use of bonding agents, if necessary
• Fiber mechanical properties, at first tensile and flexural moduli
• Fiber structure, either external, i.e., linear or branched geometry,
fibrillar, bundles, etc., or internal, i.e., hollow or plain fiber
• Fiber surface (smooth, rugged, …) and surface chemistry

Except for synthetic fibers, not all such information are easily available, and
as a matter of fact are missing for most natural fibers even for those which
have achieved a significant industrial success (e.g., Santoweb® brand). Such
a ­situation obviously limits scientific investigation on such systems, namely
considerations based on well established micromechanic models for (syn-
thetic) fiber filled thermoplastic systems. What is nevertheless well estab-
lished in what short fiber-filled rubber composites are concerned is that an
average length (or length-to-diameter ratio) above a critical value must be kept
to achieve an effective polymer-to-fiber load transfer, plus an excellent bond-
ing between fibers and polymer matrix. It seems that, where treated cellulose
fibers are concerned, this is effectively the case because the minimum critical
aspect ratio is relatively low and silica-resorcinol-­hexamethylenetetramine
systems appear as optimal in terms of cost/performance ratio. Murty and
De86 on Natural Rubber (NR)/jute systems, and Chakraborty et al.87 on car-
boxylated nitrile rubber (XNBR)/jute composites found for instance that an
aspect ratio as low as forty still gives a good reinforcement, likely because an
excellent bonding is achieved between the fibers and the rubber matrix. Fiber
orientation is also an important aspect and it is well understood that below an
aspect ratio of around 100, and providing the fiber content is not excessively
high to avoid bundling, beneficial orientation effects can be obtained through
adequate processing (as previously explained). Such considerations allow
Polymers and Short Fibers 383

to somewhat understand why timing V-belts and water coolant hoses have
become quite traditional applications for treated cellulose–rubber systems.

References
1. W. Voigt. Über die Beziehung zwischen den beiden Elastizitätskonstanten iso-
troper Körper. Wiedemanns Ann. Phys. u. Chem. (Leipzig), 38, 573–587, 1889.
2. A. Reuss. Berechnung der Fliessgrenzen von Mischkristalle auf Grund der
Plastizitätsbedingung für Einskristalle. Z. angew. Math. u. Mech., 9, 49–58, 1929.
3. S. Abrate. The mechanics of short-fiber reinforced composites: a review. Rubb.
Chem. Technol., 59, 384–404, 1986.
4. C.L. Tucker III, E. Liang. Stiffness predictions for unidirectional short-fiber com-
posites : review and evaluation. Compos. Sci. Technol., 59, 655–671, 1999.
5. L.E. Nielsen. Mechanical Properties of Polymers and Composites, 2nd Ed. M. Dekker,
New York, NY, 1994. ISBN 0-8247-8964-4.
6. B.D. Agarwal, L.J. Broutman. Analysis and Performance of Fiber Composites, 2nd
Ed. Wiley-Interscience, New York, NY, 1990. ISBN-10: 0471511528.
7. S.K. De, J.R. White, Eds. Short-fibre Polymer Composites. Woodhead, Cambridge,
UK, 1996. ISBN: 1-85573 230-3.
8. B.W. Rosen, N.F. Dow. Mechanics of failure of fibrous composites. In Fracture,
An Advanced Treatise, Volume 7, H. Liebowitz Ed. Academic Press, New York,
NY, 612–672, 1972.
9. J.C. Halpin. Stiffness and expansion estimates for oriented short fiber compos-
ites. J. Compos. Mater., 3, 732–734, 1969.
10. J.C. Halpin, J.L. Kardos. The Halpin-Tsai equations: a review. Polym. Eng. Sci.,
16, 344–352, 1976.
11. R.L. Hewitt, M. de Malherbe. An approximation for the longitudinal shear mod-
ulus of continuous fibre composites. J. Compos. Mater., 4, 280–282, 1970.
12. J.L. Kardos. Critical issues in achieving desirable mechanical properties for
short fiber composites. Pure Appl. Chem., 57 (11), 1651–1657, 1985.
13. L.A. Goettler. Mechanical property enhancement in short-fiber composites
through the control of fiber orientation during fabrication. Polymer Comp., 5,
60–71, 1984.
14. J.L. Leblanc, G. Cervantès, B. Lisiecki. Estimating the mean orientation of
short glass fibres in injected thermoplastic parts from stiffness measurements.
Proceedings EURO-FILLERS 95 Conference, Mulhouse, France, Sept. 11–14, 413–
418, 1995.
15. L.E. Nielsen. Generalized equation for the elastic moduli of composite ­materials.
J. Appl. Phys., 41, 4626–4627, 1970.
16. T.B. Lewis, L.E. Nielsen. Dynamic mechanical properties of particulate-filled
composites. J. Appl. Polym. Sci., 14, 1449–1471, 1970.
17. T. Mori, K. Tanaka. Average stress in matrix and average elastic energy of mate-
rials with misfitting inclusions. Acta Metallurgica, 21, 571–574, 1973.
18. Y. Benveniste. A new approach to the application of Mori-Tanaka’s theory in
composite materials. Mech. Mater., 6, 147–157, 1987.
384 Filled Polymers

19. G.P. Tandon, G.J. Weng. The effect of the aspect ratio of inclusions on the elastic
properties of unidirectionally aligned composites. Polym. Compos., 5, 327–333, 1984.
20. J.D. Eshelby. The determination of the elastic field of an ellipsoidal inclusion,
and related problems. Proc. Royal Soc., A241, 376–396, 1957.
21. Z. Gao, A.H. Tsou. Mechanical properties of polymers containing fillers. J.
Polym. Sci., B: Polym. Phys., 37, 155–172, 1999.
22. H.L. Cox. The elasticity and strength of paper and other fibrous materials. Brit.
J. Appl. Phys., 3, 72–79, 1952.
23. B.W. Rosen. Tensile failure of fibrous composites. Am. Instit. Aeronautics
Astronautics J., 2, 1985–1991, 1964.
24. G.P. Karman, K.L. Reifsnider. Micromechanics of short-fiber composites. Compos.
Sci. Technol., 43, 137–146, 1992.
25. V.I. Räisänen, M.J. Alava, K.J. Niskanen, R.M. Nieminen. Does the shear-lag
model apply to random fiber networks? J. Mater. Res., 12 (10), 2725–2732,
1997.
26. P.S. Theocaris. Mesophase concept in composites. In Polymers—Properties and
Applications. Springer Series. Springer Verlag, New York, NY, 1987. ISBN-10:
0387150528.
27. L. Czarnecki, J.L. White. Shear flow rheological properties, fiber damage, and
mastication characteristics of aramid-, glass-, and cellulose-fiber-reinforced
polystyrene melts. J. Appl. Polym. Sci., 25, 1217–1244, 1980.
28. G. Kalaprasad, G. Mathew, C. Pavithran, Sabu Thomas. Melt rheological behav-
ior of intimately mixed short sisal-glass hybrid fiber-reinforced low-density
polyethylene composites. I. Untreated fibers. J. Appl. Polym. Sci., 89, 432–442,
2003.
29. R. Guo, J. Azaiez, C. Bellehumeur. Rheology of fiber filled polymer melts: Role
of fiber-fiber interactions and polymer-fiber coupling. Polym. Eng. Sci., 45, 385–
399, 2005.
30. H.M. Laun. Orienation effects and rheology of short glas fiber-reinforced ther-
moplastics. Colloid Polym. Sci., 262, 257–269, 1984.
31. R.J. Crowson, M.J. Folkes, P.F. Bright. Rheology of short glass fiber-reinforced
thermoplastics and its application to injection molding I. Fiber motion and vis-
cosity measurement. Polym. Eng. Sci., 20, 925–933, 1980.
32. A.T. Mutel, M.R. Kamal. The effect of glass fibers on the rheological behavior
of polypropylene melts between rotating parallel plates. Polym. Compos., 5 (1),
29–35, 1984.
33. A.T. Mutel, M.R. Kamal. Characterization of the rheological behavior of fiber-
filled polypropylene melts under steady and oscillatory shear using cone-and-
plate and rotational parallel plate rheometry. Polym. Compos., 7 (5), 283–294,
1986.
34. Y. Chan, J.L. White, Y. Oyanagi. Influence of glass fibers on the extrusion and
injection molding characteristics of polyethylene and polystyrene melts. Polym.
Eng. Sci., 18 (4), 268–272, 1978.
35. Y. Chan, J.L. White, Y. Oyanagi . A fundamental study of the rheological proper-
ties of glass fibers filled polyethylene and polystyrene melts. J. Rheol., 22, 507–
524, 1978.
36. B.A. Knutsson, J.L. White, K.B. Abbas. Rheological and extrusion characteris-
tics of glass fiber-reinforced polycarbonate. J. Appl. Polym. Sci., 26, 2347–2362,
1981.
Polymers and Short Fibers 385

37. R.J. Crowson, M.J. Folkes. Rheology of short glass fiber-reinforced thermoplas-
tics and its application to injection molding. II. The effect of material param-
eters. Polym. Eng. Sci., 20, 934–940, 1980.
38. S. Kenig. Fiber orientation development in molding of polymer composites.
Polym. Compos., 7 (1), 50–55, 1986.
39. M. Vincent, J.F. Agassant. Experimental study and calculations of short glass
fiber orientation in a center gated molded disc. Polym. Compos., 7 (2), 76–83,
1986.
40. Å. Larsen. Injection molding of short fiber reinforced thermoplastics in a center-
gated mold. Polym. Compos., 21 (1), 51–64, 2000.
41. A.T. Mutel, M.R. Kamal. Rheological behavior and fiber orientation in slit flow
of fiber reinforced thermoplastics. Polym. Compos., 12 (3), 137–145, 1991.
42. G.B. Jeffery. The motion of ellipsoidal particles immersed in a viscous fluid.
Proc. Royal Soc. London, A102, 161–179, 1922.
43. R. Takserman-Krozer, A. Ziabicki. Behaviour of polymer solutions in a velocity
field with parallel gradient. I. Orientation of rigid ellipsoids in a dilute solution.
J. Polym. Sci. A, 1, 491–506, 1963.
44. Y. Iso, D.L. Koch, C. Cohen. Orientation in simple shear flow of semi-dilute
fiber suspensions 1. Weakly elastic fluids. J. Non-Newtonian Fluid Mech., 62 (2–3),
115–134, 1996.
45. Y. Iso, C. Cohen, D.L. Koch. Orientation in simple shear flow of semi-dilute
fiber suspensions 2. Highly elastic fluids. J. Non-Newtonian Fluid Mech., 62 (2–3),
135–153, 1996.
46. A. Bledzki, J. Gassan. Composites reinforced with cellulose based fibers. Progr.
Polym. Sci., 24, 221–274, 1999.
47. R.M. Rowell, A. Tillman, R. Simonson. A simplified procedure for the acetyla-
tion of hardwood and softwood flakes for flakeboard production. J. Wood Chem.
Technol., 6, 427–448, 1986.
48. P.S. Mukhersee, K.G Satyanarayana. An empirical evaluation of structure-prop-
erty relationships in natural fibres and their fracture behaviour. J. Mater. Sci., 21,
4162–4168, 1986.
49. C. Pavithran, P.S. Mukherjee, M. Brahmakumar, A.D. Damodaran . Impact
properties of natural fiber composites. J. Mater. Sci. Lett., 6, 882–884, 1987.
50. I. Baroulaki, J.A. Mergos, G. Pappa, P. A. Tarantili, D. Economides, K. Magoulas,
C.T. Dervos. Performance of polyolefin composites containing recycled paper
fibers. Polym. Adv. Technol., 17, 954–966, 2006.
51. D. Basu, A.N. Banerges, A. Misra. Comparative rheological studies on Jute-Fiber
and Glass-Fiber-filled Polypropylene composites melts. J. Appl. Polym. Sci., 46,
1999–2002, 1992.
52. F.M. Felix, P. Gatenholm. The nature of adhesion in composites of modified cel-
lulose fibers and polypropylene. J. Appl. Polym. Sci., 42, 620–609, 1991.
53. A.Y. Kharade, D.D. Kale. Lignin-filled polyolefins. J. Appl. Polym. Sci., 72, 1321–
1326, 1999.
54. B.V. Kokta, R.G. Raj, C. Daneault. Use of wood flour as a filler in polypropyl-
ene: Studies on mechanical properties. Polym.-Plast. Technol. Eng., 28, 247–259,
1989.
55. L.M. Matuana, R.T. Woodhams, J.J. Balatinecz, C.B. Park. Influence of interfa-
cial interactions on the properties of PVC/cellulosic fiber composites. Polym.
Compos., 19, 446–455, 1998.
386 Filled Polymers

56. J.C. Caraschi, A.Lopes Leão. Woodflour as reinforcement of polypropylene.


Mat. Res., São Carlos, Brazil, Oct./Dec.5 (4), 405–409, 2002.
57. J.Z. Lu, Q. Wu, H.S. McNabb Jr. Chemical coupling in wood fiber and polymer
composites: a review of coupling agents and treatments. Wood Fiber Sci., 32 (1),
88–104, 2000.
58. M. Kazayawoko, J.J. Balatinecz, R.T. Woodhams. Diffuse reflectance Fourier
transform infrared spectra of wood fibers treated with maleated polypropyl-
enes. J. Appl. Polym. Sci., 66,1163–1173, 1997.
59. L.M. Matuana, J.J. Balatinecz, R.N.S. Sodhi, C.B. Park. Surface characterization
of esterified cellulose fibers by XPS and FTIR spectroscopy. Wood Sci. Technol.,
35, 191–201, 2001.
60. N.M. Stark, R.E. Rowlands. Effects of wood fiber characteristics on mechanical
properties of wood/polypropylene composites. Wood Fiber Sci., 35 (2), 167–174,
2003.
61. L. Dányádi, K. Renner, Z. Szabó, G. Nagy, J. Móczó, B. Pukánszky. Wood flour
filled PP composites: adhesion, deformation, failure. Polym. Adv. Technol., 17,
967–974, 2006.
62. L. Dányádi, K. Renner , J. Móczó, B. Pukánszky. Wood flour filled polypropyl-
ene composites: Interfacial adhesion and micromechanical deformations. Polym.
Eng. Sci., 47, 1246–1255, 2007.
63. R.C.R. Nunes, J.E.S. Affonso. Interaction of NBR rubber with regenerated cel-
lulose. Kautsch., Gummi Kunstst., 52, 787–798, 1999.
64. A.F. Martins, L.L.Y. Visconte, R.C.R. Nunes. Evaluation of natural rubber and
cellulose II compositions by curing and mechanical properties. Kautsch., Gummi
Kunstst., 55, 637–641, 2002.
65. K. Brandt, R.H. Schuster, R.C.R. Nunes. Influence of process parameters on the
viscoelastic behavior of NBR-Cellulose II nanocomposites mixtures. Kautsch.,
Gummi Kunstst., 59, 511–515, 2006.
66. A.Y. Coran, K. Boustany, P. Hamed. Unidirectional fiber-polymer composites:
swelling and modulus anisotropy. J. Appl. Polymer Sci., 15, 2471– 2485, 1971.
67. A.Y. Coran, K. Boustany, P. Hamed. Short-fiber-rubber composites : the proper-
ties of oriented cellulose-fiber-elastomer composites. Rubb. Chem. Technol., 47,
396–410, 1974.
68. A.Y. Coran, P. Hamed, L.A. Goettler. The mechanical behavior of short-fiber
elastomer composites. Rubb. Chem. Technol., 49, 1167–1181, 1976.
69. K. Boustany, A.Y. Coran (to Monsanto Co). Preparation of discontinuous fiber
reinforced elastomer. U.S. Patent 3,836,412, 1974.
70. L.A. Goettler, K.S. Shen. Short fibers reinforced elastomers. Rubb. Chem. Technol.,
56, 619–638, 1983.
71. R.S. Rajeev. Fibers-reinforced elastomers. In Current Topics in Elastomers Research,
A.K. Bhowmick Ed. CRC Press, Taylor & Francis Group, Boca Raton, FL, 351–
394, 2008. ISBN-13: 978-0-8493-7317-6.
72. E. Morita. Reactions of resin formers in dry bonding rubber systems. Rubb.
Chem. Technol., 53, 795–804, 1980.
73. J.E. O‘Connor. Short-fiber-reinforced elastomer composites. Rubb. Chem. Technol.,
50, 945–958, 1977.
74. D.K. Setua, B. Dutta. Short silk fiber-reinforced polychloroprene rubber com-
posites. J. Appl. Polym. Sci., 29, 3097–3114, 1984.
Polymers and Short Fibers 387

75. L. Ibarra Rueda, C. Chamorro Antón, M.C. Tabernero Rodriguez. Mechanics of


short fibers in filled styrene-butadiene rubber (SBR) composites. Polym. Compos.,
9 (3), 198–203, 1988.
76. C. Vajrasthira, T. Amornsakchai, S. B.-Limcharoen. Fiber-matrix interactions in
aramid-short-fiber-reinforced thermoplastic polyurethane composites. J. Appl.
Polym. Sci., 87, 1059–1067, 2003.
77. W.B. Wennekes, R.N. Datta, J.W. Noordermeer, F. Elkink. Fiber adhesion to rub-
ber compounds. Rubb. Chem. Technol., 81, 523–540, 2008.
78. A.Y. Coran, K. Boustany, P. Hamed. Unidirectional fiber-polymer composites:
swelling and modulus anisotropy. J. Appl. Polymer Sci., 15, 2471, 1971.
79. W. Guo, M. Ashida. Mechanical properties of PET short fiber-polyester thermo-
plastic elastomer composites. J. Appl. Polym. Sci., 49, 1081–1091, 1993.
80. R.S. Rajeev, A.K. Bhowmick, S.K. De, G.J.P. Kao, S. Bandyopadhyay. New com-
posites based on short melamine fiber reinforced EPDM rubber. Polym. Compos.,
23 (4), 574–591, 2002.
81. L.A. Goettler, A.J. Lambright (to Monsanto Co). Process for controlling orienta-
tion of discontinuous finer in a fiber reinforced product formed by extrusion.
U.S. Patent 4,056,591, 1977.
82. L.A. Goettler, A.J. Lambright, R.I. Leib, P.J. DiMauro. Extrusion-shaping of
curved hose reinforced with short cellulose fibers. Rubb. Chem. Technol., 54, 277–
301, 1981.
83. L.A. Goettler, J. Sezna, P.J. DiMauro. Short fiber reinforcement of extruded rub-
ber profiles. Rubb. World, 187, 33–42, 1982.
84. L.A. Goettler. The extrusion and performance of plasticized Poly(viny1 chlo-
ride) hose reinforced with short cellulose fibers. Polym. Compos., 4 (4), 249–255,
1983.
85. L.A. Goettler (to Monsanto Co). Treated fibers and bonded composites of cel-
lulose fibers in vinyl chloride polymer characterized by an isocyanate bonding
agent. U.S. Patent 4,376,144, 1981.
86. V.M. Murty and S.K. De. Short jute fiber reinforced rubber composites. Rubb.
Chem. Technol., 55, 287–308, 1982.
87. S.K. Chakraborty, D.K. Setua, S.K. De. Short jute fiber reinforced carboxylated
nitrile rubber. Rubb. Chem. Technol., 55, 1286–1307, 1982.
88. J.W. Rogers. The use of fibers in V-belt Compounds. Rubb. World, 183 (6), 27–31,
1981.
89. M. Ashida, T. Noguchi, S. Mashimo. Dynamic moduli for short Fiber-CR
­composites. J. Appl. Polym. Sci., 29, 661–670, 1984.
90. M. Ashida, T. Noguchi, S. Mashimo. Effect of matrix's type on the dynamic
properties for short fiber-elastomer composite. J. Appl. Polym. Sci., 30, 1011–1021,
1985.
91. L. Ibarra. Dynamic properties of short fiber-EPDM matrix composites as a func-
tion of strain amplitude. J. Appl. Polym. Sci., 54, 1721–1730, 1994.
92. R.S. Rajeev, S.K. De, A.K. Bhowmick, G.J.P. Kao, S. Bandyopadhyay. Atomic
force microscopy studies of short melamine fiber reinforced EPDM rubber.
J. Mater. Sci., 36, 2621–2632, 2001.
93. S. Jin, Y. Zheng, G. Gao, Z. Jin. Effect of polyacrylonitrile (PAN) short fiber on
the mechanical properties of PAN/EPDM thermal insulating composites. Mater.
Sci. Eng. A, 483–484, 322–324, 2008.
388 Filled Polymers

94. N.C. Das, T.K. Chaki, D. Khastgir, A. Electromagnetic interference shielding


effectiveness of ethylene vinyl acetate based conductive composites containing
carbon fillers. Chakraborty. J. Appl. Polym. Sci., 80, 1601–1608, 2001.
95. S.S. Bhagawan, D.K. Tripathy, S.K. De. Stress relaxation in short jute fiber-
­reinforced nitrile rubber composites. J. Appl. Polym. Sci., 33, 1623–1639, 1987.
96. V.M. Murty, S.K. De. Effect of particulate fillers on short jute fiber-reinforced
natural rubber composites. J. Appl. Polym. Sci., 27, 4611–4622, 1982
97. V.M. Murty, S.K. De, S.S. Bhagawan, R. Sivaramakrishnan, S.K. Athithan.
Viscoelastic properties of short-fiber-reinforced rubber composites and the role
of adhesion. J. Appl. Polym. Sci., 28, 3485–3495, 1983.
98. D.K. Setua, S.K. De. Short silk fiber-reinforced Natural Rubber composites.
Rubb. Chem. Technol., 56, 808–826, 1983.
99. A.F. Younan, M.N. Ismail, A.I. Khalaf. Thermal stability of natural rubber-
­polyester short fiber composites. Polym. Degradation Stability, 48, 103–109, 1995.
100. N. Kikuchi. Tires made of short fiber reinforced rubber. Rubb. World, 214 (3),
31–32, 1996.
101. T.D. Sreeja. Cure characteristics and mechanical properties of Natural Rubber—
short Nylon fiber composites. J. Elast. Plast., 33 (3), 225–238, 2001.
102. D. De, D. De, B. Adhikari. The effect of grass fiber filler on curing characteris-
tics and mechanical properties of natural rubber. Polym. Adv. Technol., 15 (12),
708–715, 2004.
103. E. Osabohien, S.H.O. Egboh. Utilization of bowstring hemp fiber as a filler in
natural rubber compounds. J. Appl. Polym. Sci., 107, 210–214, 2008.
104. M.A. López Manchado, M. Arroyo. Short fibers as reinforcement of rubber
­compounds. Polym. Compos., 23 (4), 666–673, 2002.
105. H. Anuar, W.N. Wan Busu, S.H. Ahmad, R. Rasid. Reinforced thermoplastic
Natural Rubber hybrid composites with Hibiscus Cannabinus, L and short glass
fiber—part I: processing parameters and tensile properties. J. Compos. Mater., 42,
1075–1087, 2008.
106. M. Epstein, R.L. Shishoo. Measurement of adhesion between fibers and fast cur-
ing elastomer resin. J. Appl. Polym. Sci., 50, 863–874, 1993.
107 V.M. Murty, S.K. De. De. Short-fiber-reinforced styrene-butadiene rubber
­composites. J. Appl. Polym. Sci., 29, 1355–1368, 1984.
108 L. Ibarra Rueda, C. Chamorro Antón, M.C. Tabernero Rodriguez. Short-fiber-
reinforced styrene-butadiene rubber composites. Polym. Compos., 9 (3), 198–203,
1988.
109. R.P. Kumar, M.L.G. Amma, S. Thomas. Short sisal fiber reinforced styrene-
butadiene rubber composites. J. Appl. Polym. Sci., 58, 597–612, 1995.
Polymers and Short Fibers 389

Appendix 7

A7.1 Short Fiber-Reinforced Composites:


Minimum Fiber Aspect Ratio
The shear-lag analysis of fiber-reinforced composites considers that below a
certain length, a fiber is no longer effective in supporting the transfer of load
from the matrix. There is thus a minimum effective fiber length Lmin that can
be assessed as :

D: fiber diameter
1 Efib 1 − Φ fib Efib: tensile modulus of the fiber
Lmin = D ⋅ ⋅ ⋅
2 Gmat Φ fib Gmat: shear modulus of the matrix
Φfib: fiber volume fraction
Numerical illustration:

D: = 0.5 · 10−6·m Φfib: = 0,0.01..0.50   : fiber volume fraction range


Efib: = 73 · 109·Pa  = 73 GPa: typical tensile modulus for E-glass fiber
Emat: = 3 · 10 ·Pa
9  = 3 GPa: tensile modulus range for poly amide 66

For most thermoplastics, the shear modulus is consideraly smaller than the
tensile modulus, let’s say 1000 times smaller:

Emat
Gmat : =    Gmat = 3 · 106 ·Pa
1000

A7.1.1 Effect of Volume Fraction on Effective Fiber Length

1 . Efib 1 − Φ fib Efib


Lmin ( Φ fib ) : = D ⋅ ⋅    Fiber/matrix moduli ratio = 2.433 ⋅ 10 4
2 Gmat Φ fib Gmat
390 Filled Polymers

2.10–4
Effective fiber length, m

1.5.10–4

1.10–4

5.10–5
Above 10% fibers loading,
the effective fiber length
0
0 0.2 0.4 0.6 slightly decreases with
Fiber volume fraction increasing fibers content

Lmin ( 0.15 )
Lmin ( 0.15 ) = 6.937 ⋅ 10−5 m = 138.735
D
Lmin ( 0.20 )
Lmin ( 0.20 ) = 6.132 ⋅ 10−5 m = 122.633
D
Lmin ( 0.30 )
Lmin ( 0.30 ) = 5.012 ⋅ 10−5 m = 100.232
D

A7.1.2 Effect of Matrix Modulus on Effective Fiber Length


Gmat: = 0.1 ⋅ 106 ⋅ Pa, 0.2 ⋅ 106 ⋅ Pa.3 ⋅ 106 ⋅ Pa   Constant fiber loading: Φfib: = 0.2

1 Efib 1 − Φ fib
Lmin (Gmat ) : = D ⋅ ⋅ ⋅
2 Gmat Φ fib

4.10–4
Effective fiber length, m

2.10–4

0
0 5.105 1.106 15.106 2.106 2.5 .106 3.106
Shear modulus of matrix, Pa

Polymers and Short Fibers 391

i: = 0 .. 3 Gmat i : = Efib


= Lmin (Gmat i )m = Lmin (Gmat i )
Gmat i =
D
7.3 · 106 · Pa
1 · 104 3.931 · 10−5 78.615
7.3 · 10  · Pa
5

1 · 105 1.243 · 10 −4 248.603
7.3 · 104 · Pa
1 · 106 3.931 · 10−4 786.151
7.3 · 103 · Pa
1 · 107 1.243 · 10−3 2.486 · 103

A7.1.3 Effect of Fiber-to-Matrix Modulus Ratio on


Effective Fiber Length/Diameter Ratio
EGrat: = 0 .. 2.105    Φfib: = 0.2

1 1 − Φ fib
LD min ( EG rat ) := ⋅ EG rat ⋅
2 Φ fib

400

4 . 104
Fiber length/diameter ratio

200
   ⇐ LD min ( 4 ⋅ 10 4 ) = 157.23

0
0 5 . 104 1 . 105 1.5 . 105 2 . 105
Fiber-to-matrix modulus ratio

A7.2 Halpin–Tsai Equations for Short Fibers


Filled Systems: Numerical Illustration
Materials’ data:  µm: = 10−6 ⋅ m  nm: = 10−9 ⋅ m  GPa : = 109 ⋅ Pa : defining units

Polymer matrix (e.g., PP):

Emat
Emat: = 1.14 ⋅ GPa   νmat: = 0.43  Gmat : =   Gmat: = 0.399 °GPa
2 ( 1 + νmat )
392 Filled Polymers

Fibers (e.g., E-glass; typical diameter 10 μm):

Efib
Efib: = 77.0⋅GPa   νfib: = 0.20   Gfib : =    Gfib: = 32.083 °GPa
2 ( 1 + νfib )

L
L: = 1 ⋅ mm   D: = 5 ⋅ 10−3 ⋅ mm   α : = : aspect ratio    α = 200
D

Flexural modulus (GPa) for


Φ: = 0, 0.01 .. 1  : fiber volume ratio range short glass fibers filled PP
Average data various suppliers
A7.2.1 Longitudinal (Tensile) Modulus E 11
Flex
L Φ
ζL(Φ) := 2 ⋅  Efib  SGF Mod
D  E  − 1 % wt % wt
GPa
µ(Φ) := mat
 Efib  0 1.152 
 E  + ζL(Φ)  10
0

mat
 0.0337 1.87 
 15 0.05405 3.19 
 
20 0.06865 3.34 
Data: = 
30 0.1017 5.36 
 1 + ζL ( Φ ) ⋅ µ(Φ) ⋅ Φ   
Ex cpd (Φ) := Emat ⋅    40 0.1318 7.1 
 1 − µ (Φ ) ⋅ Φ  
 50 0.1605 8.964 
 60 0.1952 18.07 

100 30
Longitudinal/flexural modulus, GPa

Longitudinal/flexural modulus, GPa

0.9

20

50

10

0 0
0 0.5 1 0 0.1 0.2 0.3
Fiber volume fraction Fiber volume fraction
Model PP+SGF Model PP+SGF

The Halpin–Tsai equation predicts a linear variation of the longitudinal


­modulus with increasing fiber volume fraction; such a behavior is not exhib-
ited by commercial PP + SGF composites likely because fibers are far to be
perfectly aligned in such materials.
Polymers and Short Fibers 393

A7.2.2 Transversal (Tensile) A7.2.3 Shear Modulus G12


Modulus E 22

ζT(Φ): = 2 + 40 ⋅ Φ10 ζG(Φ): = 1 + 40 ⋅ Φ10

40

π
2. 3

20

0
0 0.5 1
Φ
ζT(Φ) ζG(Φ)

 0.1   2   1 
 0.2   2   1 
     
 0.3   2   1  For fiber fractions
     
φ : =  0.4    ζT(φ) =  2.004    ζG(φ) =  1.004  below 0.4, the 40 Φ10
 0.5   2.039   1.039  term is negligible
     
 0.6   2.242  1.242 
 0.7   3.13   2.13 
     

 Efib   Gfib 
 E  − 1  G  − 1
µT (Φ) := mat
µG(Φ) : = mat
 Efib   Gfib 
 E  + ζT (Φ)  G  + ζG(Φ)
mat mat

 1 + ζT (Φ) ⋅ µT (Φ) ⋅ Φ   1 + ζG(Φ) ⋅ µG(Φ) ⋅ Φ 


Ey cpd (Φ) : = Emat ⋅   Gcpd (Φ) := Gmat ⋅  
 1 − µT (Φ) ⋅ Φ  1 − µG(Φ) ⋅ Φ
394 Filled Polymers

100
40
Transverse modulus, GPa

0.9 0.9

Shear modulus, GPa


50 20

0 0
0 0.5 1 0 0.5 1
Fiber volume fraction Fiber volume fraction

A7.2.4 Modulus for Random A7.2.5 Fiber Orientation as an


Fiber Orientation Adjustable Parameter
x : = 0.560
3 5
Ercpd (Φ) := ⋅ Ex cpd (Φ) + ⋅ Ey cpd (Φ)
8 8 Eracpd(Φ) : = x ⋅ Excpd(Φ) + (1−x) ⋅ Eycpd(Φ)

30 30

20 20
Modulus, GPa
Modulus, GPa

10 10

0 0
0 0.1 0.2 0 0.1 0.2
Fiber volume fraction Fiber volume fraction
Model Ex Model Ex PP+SGF
PP+SGF
Model Ey Random orientation Model Ey 0.56

A7.2.6 Average Orientation Parameters from Halpin–Tsai


Equations for Short Fibers Filled Systems
Materials’ data:  µm: = 10−6 ⋅ m  nm: = 10−9 ⋅ m  GPa: = 109 ⋅  Pa: defining units

L
SGF: Efib: = 73.0 ⋅ GPa   L: = 300 ⋅ µm  D: = 10 ⋅ µm  α: = : aspect ratio   α = 30
D
Polymers and Short Fibers 395

Polymer matrix and Composites:


Injection molded fatigue test samples (ASTM D1708)
Tensile measurements (ASTMD638) on specimens (3 × 5 × 10 mm) cut out
of fatigue samples

PBT + SGF Φ Fiber Emeas PA/PAT+SGF Φ Fiber Emeas


GPa GPa

 0 2.6   0 3.2 
 0.115  0.1308 8.431 
6.475    
Data1: =  Data2: = 
 0.1735 8.445   0.1938 10.771 
   
 0.3288 14.265   0.2962 14.930 

E1mat : = ( Data1<1> )0 ⋅ GPa E2 mat : = ( Data2 <1> )0 ⋅ GPa

Φ1 : = Data1< 0> Φ2 : = Data2 < 0>

i: = 0..3  E1meas. : = ( Data1<1> )i ⋅ GPa E2 meas. : = ( Data2 <1> )i ⋅ GPa


i i

A7.2.6.1  Longitudinal (Tensile) Modulus E11

L
ζL: = 2 ⋅
D

 Efib   Efib 
 E1  − 1  E2  − 1
µL1: = mat
µL2: = mat
 Efib   Efib 
 E1  + ζL  E2  + ζL   µL2 = 0.263
mat   µL1 = 0.307    mat

1 + ζL ⋅ µL1 ⋅ Φ1i 1 + ζL ⋅ µL2 ⋅ Φ 2 i


Ex cpd1i : = E1mat ⋅ Ex cpd2i : = E2 mat ⋅
1 − µL1 ⋅ Φ1i 1 − µL2 ⋅ Φ 2 i
396 Filled Polymers

A7.2.6.2  Transversal (Tensile) Modulus E22


ζT: = 2

 Efib   Efib 
 E1  − 1  E2  − 1
µT 1: = mat
µT 2: = mat
 Efib   Efib 
 E1  + ζT  E2  + ζT
mat mat

 1 + ζT ⋅ µT 1 ⋅ Φ1i   1 + ζT ⋅ µT 2 ⋅ Φ 2 i 
Ey cpd1i : = E1mat ⋅  Ey cpd2i : = E2 mat ⋅ 
 1 − µT 1 ⋅ Φ1i   1 − µT 2 ⋅ Φ 2 i 

A7.2.6.3  Orientation Parameter X

E1measi − Ey cpd1i E2 measi − Ey cpd2i


X 1i : = X2 i : =
Ex cpd1i − Ey cpd1i Ex cpd2i − Ey cpd2i

0 0.8 0
Orientation parameter X

0.606 0.697
X1 = X2 =
0.587 0.657
0.6
0.576 0.636

0.4
0 0.2 0.4
Fiber volume fraction
PBT+SGF
PA/PAT+SGF

A7.3 Nielsen Modification of Halpin–Tsai Equations with


Respect to the Maximum Packing Fraction:
Numerical Illustration
Materials’ data:  µm: = 10−6⋅m  nm: = 10−9⋅m  GPa: = 109 ⋅ Pa:defining units
Polymer matrix (e.g. PP):
Emat
Emat: = 1.14⋅GPa   νmat: = 0.43   Gmat : =    Gmat: = 0.399 °GPa
2 ( 1 + νmat )
Polymers and Short Fibers 397

Fibers (e.g. E-glass; typical diameter 10 μm):

Efib
Efib: = 77.0⋅GPa   νfib: = 0.20   Gfib : =    Gfib: = 32.083 °GPa
2 ( 1 + νfib )

L
L: = 1⋅mm   D: = 8 ⋅ 10−3 ⋅ mm   α : = : aspect ratio    α = 125
D

Φmax: = 0.6
Flexural modulus (GPa) for
short glass fibers filled PP
Φ: = 0,0.01..Φmax :fiber volume ratio range Average data various suppliers

Flex
SGE Φ Mod
A7.3.1  Maximum Packing Functions
%wt %wt GPA
[Lewis and Nielsen, J. Appl. Polym.
0 0 1.152 
Sci., 14, 1449, 1970]  10
 0.0337 1.87 
 1 − Φ max   15 0.05405 3.19 
F1(Φ) : = 1 +  ⋅Φ  
 Φ max 2  Data: = 
20 0.06865 3.34 
 30 0.1017 5.36 
 
 40 0.1318 7.1 
 50 0.1605 8.964 
1   − Φ ⋅ Φ max    
F2(Φ) := ⋅  1 − exp 
Φ   Φ max − Φ    60 0.1952 18.07 

1
2

0.5
1.5

1 0
0 0.2 0.4 0 0.2 0.4
Φ Φ
F1(Φ) F 2(Φ) F1 (Φ).Φ F 2 (Φ).Φ Φ
  

Φ: = 0.75⋅Φmax  Φ2c: = Maximize (F2,Φ)  Φ2 c := 0.496 ⇐ position of maximum


of F 2 function
398 Filled Polymers

A7.3.2 Longitudinal (Tensile) Modulus E 11


Φ: = 0,0.01…Φmax

 Efib 
 E  − 1
L
ζL := 2 ⋅ µ:= mat
D  Efib 
 E  + ζL
mat

 1 + ζL ⋅ µ ⋅ Φ   1 + ζL ⋅ µ ⋅ Φ 
Ex1cpd (Φ) := Emat ⋅  Ex2 cpd (Φ) := Emat ⋅ 
 1 − µ ⋅ F 1(Φ) ⋅ Φ   1 − µ ⋅ F 2(Φ) ⋅ Φ 

100 30
Longitudinal/flexural modulus, GPa

Longitudinal/flexural modulus, GPa

Φmax 0.12

20

50

10

0 0
0 0.5 1 0 0.1 0.2
Fiber volume fraction Fiber volume fraction
F1(phi) F2(phi) PP+SGF F1(phi) F2(phi) PP+SGF

A7.3.3 Transverse (Tensile) A7.3.4  Shear Modulus G


Modulus Ey

ζT(Φ) : = 2 + 40 ⋅ Φ10 ζG(Φ) : = 1 + 40 ⋅ Φ10

 Efib   Gfib 
 E  − 1  G  − 1
µT (Φ) := mat
µG(Φ) := mat
 Efib   Gfib 
 E  + ζT (Φ)  G  + ζG(Φ)
mat mat

 1 + ζT ( Φ ) ⋅ µT ( Φ ) .Φ   1 + ζG(Φ) ⋅ µG(Φ).Φ 
Ey cpd (Φ) : = E mat ⋅  Gcpd (Φ) : = Gmat ⋅ 
 1 − F 1( Φ ) ⋅ µT ( Φ ) ⋅ Φ   1 − F 1(Φ) ⋅ µG(Φ) ⋅ Φ 
Polymers and Short Fibers 399

100 30
Transverse modulus, GPa

Shear modulus, GPa


20

50

10

0
0 0 0.2 0.4 0.6
0 0.2 0.4 0.6
Fiber volume fraction
Fiber volume fraction

A7.4 Mori–Tanaka’s Average Stress Concept: Tandon–Weng


Expressions for Randomly Distributed Ellipsoidal
(Fiber-like) Particles: Numerical Illustration

Materials’ data:  µm: = 10−6 ⋅ m  nm: = 10−9 ⋅ m  GPa: = 109 ⋅ Pa: defining units


Polymer matrix (e.g., PP):

Emat
Emat: = 1.14 ⋅ GPa  νmat: = 0.43   Gmat : =   Gmat: = 0.399 °GPa
2 ( 1 + νmat )

Fibers (e.g., E-glass; typical diameter 10 μm):

Efib
Efib: = 77.0 ⋅ GPa   νfib: = 0.20   Gfib : =    Gfib: = 32.083 °GPa
2 ( 1 + νfib )

L
L: = 1 ⋅ mm   D: = 50 ⋅ 10−3   α : = : aspect ratio    α = 20
D

Φ: = 0,0.01..1   : fiber volume ratio range

A7.4.1 Eshelby’s Tensor (Depending on Matrix Poisson’s


Ratio and Fibers Aspect Ratio only)

α
fα : = ⋅ α ⋅ (α 2 − 1) − ac osh (α ) fα = 0.993
(α − 1)
2 3
400 Filled Polymers

1  3 ⋅ α2 − 1  3 ⋅ α2  
S1111 : = ⋅ 1 − 2 ⋅ νmat + 2 −  1 − 2 ⋅ νmat + 2 ⋅ fα 
2 ⋅ ( 1 − νmat )  α −1  α − 1  

3 ⋅ α2 1  9 
S 2222 : = + ⋅ 1 − 2 ⋅ νmat − ⋅ fα
8 ⋅ ( 1 − νmat ) ⋅ ( α 2 − 1) 4 ⋅ ( 1 − νmat )  4 ⋅ ( α 2 − 1) 

S 3333 : = S 2222

1  α2  3  
S 2233 : = ⋅ − 1 − 2 ⋅ νmat +  ⋅ fα 
4 ⋅ ( 1 − νmat )  2 ⋅ ( α − 1) 
2 4 ⋅ ( α − 1) 
2

S 3322 : = S 2233

−α 2 1  3 ⋅ α2 
S 2211 : = + ⋅ − ( 1 − 2 ⋅ νmat ) ⋅ fα
2 ⋅ ( 1 − νmat ) ⋅ ( α 2 − 1) 4 ⋅ ( 1 − νmat )  α 2 − 1 
S 3311 : = S 2211

−1  1  1  3 
S 1122 : = 1 − 2 ⋅ νmat + 2 + ⋅ 1 − 2 ⋅ νmat + ⋅ fα
2 ⋅ (1 − νmat )  α − 1  2 ⋅ (1 − νmat )  2 ⋅ (α 2 − 1) 

S1133 : = S 1122

1  α2  3  
S 2323 : = ⋅ + 1 − 2 ⋅ νmat −  ⋅ fα 
4 ⋅ ( 1 − νmat )  2 ⋅ ( α − 1) 
2 4 ⋅ ( α − 1) 
2

S 3232 : = S 2323

1  α2 + 1 1  3 ⋅ ( α 2 + 1)  
S1212 : = ⋅ 1 − 2 ⋅ νmat − 2 − ⋅ 1 − 2 ⋅ νmat −  ⋅ fα 
4 ⋅ ( 1 − νmat )  α −1 2  α −1 
2

S1313 : = S1212

A7.4.2 Materials’ Constants (i.e., Not Depending on Fiber


Volume Fraction)

Gfib − Gmat
D1 := 1 + 2 ⋅ D1 = 4.346
νfib ⋅ Efib νmat ⋅ Emat

(1 + νfib ) ⋅ (1 − 2 ⋅ νfib ) (1 + νmat ) ⋅ (1 − 2 ⋅ νmat )
Polymers and Short Fibers 401

νmat ⋅ Emat
+ 2 ⋅ Gmat
D2 : =
(1 + νmat ) ⋅ (1 − 2 ⋅ νmat ) D2 = 0.171
νfib ⋅ Efib νmat ⋅ Emat

(1 + νfib ) ⋅ (1 − 2 ⋅ νfib ) (1 + νmat ) ⋅ (1 − 2 ⋅ νmat )

νmat ⋅ Emat

D3 : =
(1 + νmat ) ⋅ (1 − 2 ⋅ νmat ) D3 = 0.129
νfib ⋅ Efib νmat ⋅ Emat

( fib ) (
1 + ν ⋅ 1 − 2 ⋅ ν fib ) ( mat ) ⋅ (1 − 2 ⋅ νmat )
1 + ν

A7.4.3 Materials and Volume Fraction Depending Constants

B1 (Φ): = Φ ⋅ D1 + D2 + ( 1 − Φ ) ⋅ ( D1 ⋅ S1111 + 2 ⋅ S2211 )

B2 (Φ) : = Φ + D3 + ( 1 − Φ ) ⋅ ( D1 ⋅ S1122 + S2222 + S2233 )

B3 (Φ) := Φ + D3 + ( 1 − Φ ) ⋅ [ S1111 + ( 1 + D1 ) ⋅ S2211 ]

B4 (Φ) := Φ ⋅ D1 + D2 + ( 1 − Φ ) ⋅ ( S1122 + D1 ⋅ S2222 + S2233 )

B5 (Φ) := Φ + D3 + ( 1 − Φ ) ⋅ ( S1122 + S2222 + D1 ⋅ S2233 )

A1 (Φ) := D1 ⋅ ( B4 (Φ) + B5 (Φ)) − 2 ⋅ B2 (Φ)

A2 (Φ) := ( 1 + D1 ) ⋅ B2 (Φ) − ( B4 (Φ) + B5 (Φ))

A3 (Φ) := B1 (Φ) − D1 ⋅ B3 (Φ)

A4 (Φ) := ( 1 + D1 ) ⋅ B1 (Φ) − 2 ⋅ B3 (Φ)

A5 (Φ) :=
(1 − D1 )

B4 (Φ) − B5 (Φ)

A(Φ) := 2. B2 (Φ) ⋅ B3 (Φ) − B1 (Φ) ⋅ ( B4 (Φ) + B5 (Φ))


402 Filled Polymers

A7.4.4  Calculating the Longitudinal (Tensile) Modulus E11

Emat
E11 (Φ) :=
( A (Φ) + 2 ⋅ νmat ⋅ A2 (Φ))
1+ Φ⋅ 1
  
E11 ( 0.2 ) = 9.286  GPa
A(Φ)

upper bound: Eup (Φ) := Emat ⋅ ( 1 − Φ ) + Efib ⋅ Φ

Fiber aspect ratio : α = 20


Emat ⋅ Efib
lower bound : Elow ( Φ ) :=
Efib ⋅ ( 1 − Φ ) + Emat ⋅ Φ

100

50
E11 is the most sensitive elastic con-
stant to the fiber aspect ratio; when
the aspect ratio becomes larger than
50, the Mori–Tanaka’s average stress
approach gives E11 not much different
from the upper bound prediction
0
0 0.5 1
Φ
E11(Φ).GPa–1
Elow(Φ).GPa–1
Eup(Φ).GPa–1

A7.4.5  Calculating the Transverse (Tensile) Modulus E22

Emat
E22 (Φ) :=
1+ Φ⋅
[ −2 ⋅ νmat ⋅ A3 (Φ) + (1 − νmat ) ⋅ A4 (Φ) + (1 + νmat ) ⋅ A5 (Φ) ⋅ A(Φ)]
2 A(Φ)

E22(0.2) = 1.907 °GPa
Polymers and Short Fibers 403

100

50

00 The transverse modulus if not really sensi-


0.5 1
Φ tive to fiber aspect ratio
E22(Φ).GPa–1
Elow(Φ).GPa–1
Eup(Φ).GPa–1

A7.4.6  Calculating the (In-Plane) Shear Modulus G12

 
 Φ 
G12 (Φ) := Gmat ⋅ 1 +  G12 ( 0.2 ) = 0.594° GPa
 Gmat
+ 2 ⋅ (1 − Φ ) ⋅ S1212 
 Gfib − Gmat 

s1212 = 0.247 upper bound: Gup (Φ) := Gmat ⋅ (1 − Φ ) + Gfib ⋅ Φ

Gmat ⋅ Gfib
lower bound : Glow (Φ) :=
Gfib ⋅ ( 1 − Φ ) + Gmat ⋅ Φ

40

30

20

10

The shear modulus if not really sensitive to


0
0 0.5 1 fiber aspect ratio
Φ
G12(Φ).GPa–1
Glow(Φ).GPa–1
Gup(Φ).GPa–1
404 Filled Polymers

A7.4.7  Calculating the (Out-Plane) Shear Modulus G23

 Φ 
1 + 
G23 (Φ) : = Gmat ⋅  G mat
+ 2 ⋅ (1 − Φ ) ⋅ S2323  G23 ( 0.2 ) = 0.572° GPa
 Gfib − Gmat 
 

s2323 = 0.28
upper bound: Gup (Φ) := Gmat ⋅ (1 − Φ ) + Gfib ⋅ Φ

Gmat ⋅ Gfib
lower bound : Glow (Φ) : =
Gfib ⋅ ( 1 − Φ ) + Gmat ⋅ Φ

40

30

20

10

0
0 0.5 1
Φ

G12(Φ).GPa–1
Glow(Φ).GPa–1
Gup(Φ).GPa–1

A7.4.8  Comparing with Experimental Data


Flexural modulus (GPa) for short glass fibers filled PP

Rem: flexural modulus = tensile modulus × (gyration radius)2


Polymers and Short Fibers 405

  Average data various suppliers Data single supplier

Flex Flex
SGF Φ Mod SGF Φ Mod
%wt % wt %wt % wt GPa
GPa
0 0 1.152  0 0 1.1303 
 10  10 2.309 
 0.0337 1.87   0.03375
 15 0.05405 3.19   13 0.04335 2.757 
  Data2: =  
20 0.06865 3.34   20 0.0656 3.654 
Data1: =   30
 30 0.1017 5.36  0.0959 5..329 
   
 400 0.1318 7.1   40 0.126 6.756 
 50 0.1605 8.964 
 
 60 0.1952 18.07 

20

10

The model meets reasonnably well


experimental data up to around 0.10
volume fraction and with a fiber ratio
0 of around 20
0 0.1 0.2
Φ, Datal<1>, Data2<1>

E11(Φ).GPa–1 Datal<2>
<2>
Datal
406 Filled Polymers

A7.4.9 Tandon–Weng Expressions for Randomly Distributed


Spherical Particles: Numerical illustration

Materials’ data:   nm: = 10−9⋅m   GPa: = 109⋅Pa   : defining units

Polymer matrix:

E mat
Emat: = 2.76⋅GPa   νmat: = 0.35   Gmat : =    Gmat = 1.022 °GPa
2 ( 1 + νmat )

Filler:

E fill
Efill: = 72.4⋅GPa   νfill: = 0.20   Gfill : =    Gfill = 30.167  °GPa
2 ( 1 + νfill )

Φ: = 0,0.01 .. 1   : filler volume ratio range

A7.4.9.1  Eshelby’s Tensor (Depending on Matrix Poisson’s Ratio Only)

7 − 5 ⋅ νmat 5 νmat ⋅ −1 4 − 5 ⋅ νmat


S1111 :=    S1122 : =    S1212 : =
15 ⋅ ( 1 − νmat ) 15 ⋅ ( mat )
1 − ν 15 ⋅ ( 1 − νmat )

s2222 : = s1111 s1313 : = s1212 s3333 : = s1111 s2323 : = s1212

s1133 : = s1122 s3322 : = s2233 s2233 : = s1122 s3232 : = s2323

s3311 : = s1122 s3131 : = s1212 s2211 : = s3311

A7.4.9.2 Materials’ Constants (i.e., Not Depending


on Filler Volume Fraction)

Gfill − Gmat
D1 : = 1 + 2 ⋅ D1 = 4.288
νfill ⋅ Efill νmat ⋅ Emat

( fill ) (
1 + ν ⋅ 1 − 2 ⋅ ν fill ) ( mat ) ⋅ (1 − 2 ⋅ νmat )
1 + ν
Polymers and Short Fibers 407

νmat ⋅ Emat
+ 2 ⋅ Gmat
D2 : =
(1 + νmat ) ⋅ (1 − 2 ⋅ νmat ) D2 = 0.25
νfill ⋅ Efill νmat ⋅ Emat

(1 + νfill ) ⋅ (1 − 2 ⋅ νfill ) (1 + νmat ) ⋅ (1 − 2 ⋅ νmat )
νmat ⋅ Emat

D3 : =
(1 + νmat ) ⋅ (1 − 2 ⋅ νmat ) D3 = 0.135
νfill ⋅ Efill νmat ⋅ Emat

( fill ) (
1 + ν ⋅ 1 − 2 ⋅ ν fill ) ( mat ) ⋅ (1 − 2 ⋅ νmat )
1 + ν

A7.4.9.3  Materials and Volume Fraction Depending Constants

B1 (Φ): = Φ ⋅ D1 + D2 + ( 1 − Φ ) ⋅ ( D1 ⋅ S1111 + 2 ⋅ S2211 )

B2 (Φ) := Φ + D3 + ( 1 − Φ ) ⋅ ( D1 ⋅ S1122 + S2222 + S2233 )

B3 (Φ) := Φ + D3 + ( 1 − Φ ) ⋅ [ S1111 + ( 1 + D1 ) ⋅ S2211 ]

B4 (Φ) := Φ ⋅ D1 + D2 + ( 1 − Φ ) ⋅ ( S1122 + D1 ⋅ S2222 + S2233 )

B5 (Φ) := Φ + D3 + ( 1 − Φ ) ⋅ ( S1122 + S2222 + D1 ⋅ S2233 )

A1 (Φ) := D1 ⋅ ( B4 (Φ) + B5 (Φ)) − 2 ⋅ B2 (Φ)

A2 (Φ) := ( 1 + D1 ) ⋅ B2 (Φ) − ( B4 (Φ) + B5 (Φ))

A3 (Φ) := B1 (Φ) − D1 ⋅ B3 (Φ)

(1 − D1 )
A4 (Φ) : = ( 1 + D1 ) ⋅ B1 (Φ) − 2 ⋅ B3 (Φ)    A5 (Φ) :=
B4 (Φ) − B5 (Φ)

A ( Φ ) := 2.B2 ( Φ ) ⋅ B3 ( Φ ) − B1 ( Φ ) ⋅ ( B4 ( Φ ) + B5 ( Φ ))
408 Filled Polymers

A7.4.9.4  Calculating the Tensile Modulus E

Emat    Guth and Gold :


E (Φ) :=
( A (Φ) + 2 ⋅ νmat ⋅ A2 (Φ))
1+ Φ⋅ 1 EGG (Φ) := Emat ⋅ (1 + 2.5 ⋅ Φ + 14.1 ⋅ Φ 2 )
A(Φ)

Emat ⋅ Efill
E ( 0.2 ) = 3.777 GPa    lower bound : Elow (Φ) :=
Efill ⋅ ( 1 − Φ ) + Emat ⋅ Φ

80

60
Tensile modulus, GPa

40

20
Rem: the longitudinal and
transverse (tensile) moduli
E11 and E22 are identical when
0
particles are spherical
0 0.5 1
Filler volume fraction
Mori-Tanaka's average stress
Lower bound modulus
Guth & Gold equation

A7.4.9.5  Calculating the Shear Modulus G

 
  Guth and Gold :
Φ
G(Φ) : = Gmat ⋅ 1 +   
 Gmat GGG (Φ) := Gmat ⋅ ( 1 + 2.5 ⋅ Φ + 14.1 ⋅ Φ 2 )
 + 2 ⋅ ( 1 − Φ ) ⋅ S1212 
Gfill − Gmat 

Gmat ⋅ Gfill
G ( 0.2 ) = 1.528 GPa   lower bound : Glow (Φ) :=
Gfill ⋅ ( 1 − Φ ) + Gmat ⋅ Φ
Polymers and Short Fibers 409

40

30
Shear modulus, GPa

20

Rem: the in-plane and out-


10
plane shear moduli G12 and
G23 are identical when par-
0 ticles are spherical
0 0.5 1
Filler volume fraction
Mori-Tanaka's average stress
Lower Bound modulus
Guth & Gold equation

A7.5  Shear Lag Model: Numerical illustration

Materials’ data:  µm: = 10−6⋅m  nm: = 10−9⋅m  GPa: = 109⋅Pa: defining units

Polymer matrix (e.g. PP):

Emat
Emat: = 1.14⋅GPa  νmat: = 0.43  Gmat : =   Gmat: = 0.399 °GPa
2 ( 1 + νmat )

Fibers (e.g. E-glass; typical diameter 10 μm):

Efib
Efib: = 77.0⋅GPa  νfib: = 0.20  Gfib : =   Gfib: = 32.083 °GPa
2 ( 1 + νfib )

L D
L: = 1⋅mm  D: = 50⋅10−3⋅mm  α : = : aspect ratio   α = 20  R : =
D 2
410 Filled Polymers

Fibers’ arrangement

π 2⋅π π π
Kr : =
2⋅ 3
  
3
[ Cox ]     2⋅ 3
[ Hexagonal ]    4
[ Cube]

Φ: = 0,0.01...09  :fiber volume ratio range Flexural modulus (GPa) for


short glass fibers filled PP
Longitudinal (tensile) modulus E11 Average data various
suppliers
Kr
  r(Φ) := R ⋅ : matrix shell radius
Φ Flex
SGF Φ Mod
%wt %wt GPa
300
Mesophase
0 0 1.152 
Mesophase thickness, microns

thickness  
r(Φ) – R:  10 0.0337 1.887 
200  15 0.05405 3.19 
 
Data: =  20 0.06865 3.34 
 30 0.1017 5.36 
 
100  40 0.1318 7.1 
 
 50 0.1605 8.964 
 60 0.1952 18.07 
0
0 0.5 1
Fiber volume fraction

2 ⋅ Gmat   β(Φ) ⋅ L  
β(Φ) :=  tanh  2  
 r(Φ)     nL (Φ) : = 1 −  : efficiency factor
R 2 ⋅ Efib ⋅ ln 
 R   β(Φ) ⋅ L 
 2 

Shear lag model equation:  E11(Φ): = nL(Φ)⋅Φ⋅Efib + (1 − Φ)⋅Emat

Shear lag model equation : 30


Longitudinal/flexural modulus, GPa

100
Longitudinal or flexural modulus, GPa

20

50
10

0
0 0.1 0.2 0.3
0 Fiber volume fraction
0 0.5 1
Fiber volume fraction

Index

A Bound rubber (BdR)


absolute value of, 116
Acetylene black, 24
concept, 108
Agglomerates
content from
voids, degree, 42
thermogravimetry analysis
Aggregates, 21, 25–26
(TGA), 118
comparison, 41
toluene extraction
diameter, size, 37
kinetic method, 117
elementary particles, assessing, 40
3D structure of, 109–110
flocculation process, 159–160
effect of
Medalia, concept of, 37–38
carbon black size on, 109–110
Medalia, occluded rubber, 42–43
elastomer and storage on, 119
solid fraction, 218
extraction kinetic method for
structure, 17
assessment, 116
Aggregates flocculation/entanglement
extraction time, 116–117
model, 218–221
factors affecting, 114–121
aggregate, solid fraction of, 219–220
kinetic aspects in formation of, 118
filler volume fraction,
and NMR results, 110–111
modulus, 220–221
50 phr Carbon Black SBR 1500
strain dependence,
Compounds, 115
modulus, 221–222
singular flow properties, 112
Alkoxy derivatives, 261
BR, see Butadiene rubber (BR)
Aluminium tri-hydrate (ATH), 263
Brunauer, Emmet, Teller (BET)
American Society for Testing Materials
method, 17, 28
(ASTM), 27, 29–30, 35
nitrogen adsorption, 238
Anchoring knots, 133
Butadiene rubber (BR), 45
Antimony tri-oxide, 263
ASTM, see American Society for
Testing Materials (ASTM) C
ATH, see Aluminium tri-hydrate (ATH) Calcite, 11
Attraction–repulsion mechanism, 154 Calcium carbonate, 285, 287–288
ground natural (GCC), 280
precipitated (PCC), 280, 290
B
Calcium oxide, 56
Bagley method, 108 white minerals used as fillers
Barrier properties, 292 for polymers, 57
BdR, see Bound rubber (BdR) Capillary rheometer, 98
BET method, see Brunauer, Emmet, Carbonates
Teller (BET) method dolomite, 56
Bingham fluid and Herschel–Bulkley GCC and PCC, 55
equation, 101 grades, (PCC), 55
Bis(triethoxysilylpropyl)tetrasulfane Carbon black (CB)
(TESPT), 236, 240 aggregates
Blanc fixe, 56 comparison, 41

411
412 Index

diameter, size, 37 level and


elementary particles, postextrusion swelling, 107
assessing, 40 temperature on dynamic
Medalia, concept of, 37–38 properties, 143
Medalia, occluded rubber, 42–43 manufacturing process for
ASTM designation vs. gas black, 24
characterization mass fractal dimension, 30, 33
data, 31–32, 35 aggregate, volume of particle, 39
characterization data, and aggregates, 33
relationships, 84 connectivity exponent, 36
data, source, 82–83 critical filler level, 43–44
DBPA of, 129 DPBA, 39–40
di-butyl phthalate (DBP) absorption Ethylene-Propylene-Diene
numbers (cDBPA), 103 Monomer rubber (EPDM)
dispersion on dynamic compounds, 38
properties, 145–148 fractal geometry, 36–37, 41
dynamic moduli, 141 fractal nature, 43
dynamic properties, 133 geometry description of
EBA/CB composites, 177 aggregates, 34
effect on type G′ and tan δ, 144 HAF grade, 40
electrical conductivity of, 176 and mean DPA absorption
fabrication processes number, 35
aggregates, 21 Medalia aggregates, concept
for furnace black, 22 of, 37–38
lamp black, manufacturing Medalia floc simulation
process, 23 approach, 38–39
methods, 22 Medalia occluded rubber, 42–43
smoke, 23–24 pellets and agglomerates, 42
thermal black process, 24 pellets mixing operations, 43
thermo-oxidative processes, 21 TEM/AIA study, 37
filled compounds void ratio, 39
extraction kinetic data on volume fraction, 41–42
N330, 117 well dispersed state, 43
morphology and nonlinear X-ray scattering experiments, 35
flow properties, 113 mechanical properties, effect on, 125
stress overshoot experiments Medalia’s floc simulation data, 85
on, 106 Mooney and Rivlin equation,
tridimensional representation 130–131
of morphology of, 111 optimum dispersion, 92–93
Filled SBR 1500 Compound, 92 particles/aggregate, 89
Gent and Park equation, 136 postextrusion swelling
Guth, Gold and Simha equation, 125 bound rubber (BdR), 109–110
interactions between rubber entrance pressure drop, 108
segments and, 169 level and, 107
inter-aggregates distances production properties and
face-centered cubic lattice process, 25
model, 93–94 properties of, 91
for optimal reinforcement, 94 on rheological properties,
junction distance, 135–136 effects of, 95
Index 413

rubber–CB interactions, 119 lamp and gas, chemical


rubber–filler interactions and, 103 functions detection, 44
rubber reinforcement by, 148–151 new equilibrium gas
shear viscosity adsorption techniques, 46–47
complex cosmetic percolation level, 46
material, 98–99 probe, 46
filled BR compound, 98 rubber–filler interactions, 45
function, effect on, 95 rubber grade, 44
Herschel–Bulkley equation, 100 surface activity, 44
plots and level, 97 ToF-SIMS and XPS, 45
power law model with Toom model, 48
yield stress, 99–100 surface energy components for, 36
silica temperature effect and tensile
comparison, low strain stress at break, 127
dynamic properties, 239 theoretical concept for, 162
comparison, surface type and temperature on
chemistry, 236 damping properties of SSBR
tensile properties, 238 compounds, 145
spatial organization and tan δ of SSBR compounds, 142
reinforcing character of, 26 usages, 21–22
specific gravity, 92 viscosity and levels of, 121–122
Stokes diameter, 93 volume fraction, 41–42, 92, 134
strain amplification factor, 131 and yield stress, 105
strain sweep tests, 178 Young’s modulus, 130
structural aspects and Carbon black pellets, 42
characterization Carbon fibers
aggregates, 25–26 aramid fibers
ASTM classification of, 29–30 polypara-phenylene
Brunauer, Emmet, Teller (BET) terephtalamide fibers, 71–72
method, 28 tensile strength and modulus, 74
cetyltrimethylammonium PAN-fibers and MPP-fibers, 71
bromide CTAB adsorption Carbon fibers reinforced epoxy
methods, 27–28 resins, 1
dibutylphthalate (DBP) Carbon nanotubes, 5
method, 28 Carboxylated nitrile rubber
DPA absorption method, 28 (XNBR), 382
elementary analysis, 26 Carcass tire applications, 30
elementary particles and Carreau equation for shear
structure of aggregates, viscosity function of
size, 28–29 polymer melts, 278
rubber reinforcement, 24–25 Carreau–Yasuda model, 268–269, 278
shapes of, 25 virgin polystyrene (PS) data,
standard characterization fitting, 324, 326–330
methods, 27 assembly result, 329–332
tread and carcass tire Casson model, 101
applications, 30 for yield stress fluids, 102
surface energy aspects Cathetometer, 96
energetic sites, 47–48 CCA, see Cluster-cluster-aggregation
esterification of, 48 (CCA) model
414 Index

cDBP absorption test Connective filaments between


maximum volume fraction, 188 rubber–filler aggregates, 112
Cellulose derivatives, 13 Connectivity exponent, 274
Cetyltriethylammonium bromide Contour length, 246
(CTAB), 17 Controlled stress rheometers, 98
adsorption methods, 27–28 Creep phenomena, 125
Characteristic time, effect of, 320 Creep resistance, 292
Chlorosilanes, 247 Critical shear rates, 269
Chlorosulfonated polyethylene Crushed dibutyl phtalate adsorption
compounds, 260 test (ASTM D-3493), 133
Chopped natural fibers, 370 CTAB, see Cetyltriethylammonium
Clay/polymer nanocomposites, 4 bromide (CTAB)
Clay(s), 300
and chemical grafting of, 50–51
D
and elastomers
grades of, 257–259 DBP, see Di-butylphthalate (DBP)
rubber extrusion and Deagglomeration-reagglomeration
calendering, 257 of filler aggregates,
exfoliated, 301 Kraus model, 155
minerals, 11 Degussa
one-step and two-step Si264 TCPTS, 240
grafting technique, 50 Si69 TESPT, 240
mechanisms, 51 De Mattia fatigue tests, 260
physical adsorption of Deutsches Institute für
functional polymers, 50 Normung (DIN), 27
Cluster-cluster-aggregation (CCA) Diamond, 265
model, 158, 218–221 Di-butylphtalate absorption (DPBA),
aggregate, solid fraction 39–40, 43
of, 219–220 Di-butyl phthalate (DBP), 17, 35, 37, 39
filler volume fraction, absorption numbers (cDBPA), 103
modulus, 220–221 method, 28
strain dependence, Diene elastomers
modulus, 221–222 carbon black and (untreated)
Cohen–Addad percolation model silica, comparison, 237–239
maximum adsorbed polymer reinforcement, 239–246
at saturation, 251 DIN, see Deutsches Institute für
for silica-PDMS, 120 Normung (DIN)
Coloring carbon blacks, 173 Dolomite, 56
Commercial filled polypropylene DPBA, see Di-butylphtalate
compounds absorption (DPBA)
volume fraction, effect of Dynamic moduli variation
mica, 300 with strain amplitude, 141–142
talc, 295–296 with temperature, 142–144
Commercial PP–mica composites Dynamic strain softening (DSS)
data, 299 fractional linear solid
Complex modulus, 255 model, 222–223
Concept of maximum volume Lion model, 167
fraction, 266 mathematical aspects, 226
Conductive carbon blacks, 175 modeling, 223–225
Index 415

Maier and Göritz model, 227–232 rubber matrix, 375–376


percolation theory, 153 short fiber reinforced rubber
Dynamic stress softening (DSS), 141 composites, 379–380
effect of mixing duration on in extruded rubber hoses,
magnitude, 148 controlling, 381
as filler network effect, 152 special fiber–elastomer
as filler–polymer network effect, 168 composites, 378–379
nonlinear viscoelastic specific gravity, 92
properties, 151 Stokes diameter, 93
origins of, 151–152 strain amplification in, 130
Payne effect, 151 styrene butadiene copolymer
stress/strain proportionality, 151 (SBR), 376–377
and talc
fractured surfaces, 260
E
ground talc, 260
Effective fiber length technical constraints, 382
effect thermoplastic systems
fiber-to-matrix modulus ratio, 391 micromechanic models for
matrix modulus, 390–391 (synthetic) fiber, 382
volume fraction, 389–390 volume fraction of, 92
Elasticity dissipation structure, 114 Elastomers and white fillers, 235
Elastic modulus, 166, 212, 223, 230 elastomers and clays, 257–260
effect of carbon black specific elastomers and silica, 235
area on, 144 diene elastomers carbon black,
temperature effect, 143 comparison, 237–239
Elastomers polydimethylsiloxane, 246–257
carbon black filled SBR compound, silanisation, reinforcement
mechanical properties, 381 diene, 239–245
cellulose fibers and chopped surface chemistry of, 235–237
aramid fibers industrial elastomers and talc, 260–266
importance, 380–381 Engineering plastics, 301
extrusion moving die ENR, see Epoxidized natural
technology, 380 rubber (ENR)
fiber orientation, 382–383 EPDM, see Ethylene-Propylene-Diene
glass fibers exhibit, 381–382 Monomer rubber (EPDM)
hexa(methoxymethyl)melamine compounds
(HMMM), 377 Epoxidized natural rubber (ENR), 45
inter-aggregates distances EPR, see Ethylene-Propylene
face-centered cubic lattice rubber (EPR)
model, 93–94 Equivalent circuit model, 172
for optimal reinforcement, 94 Eshelby’s tensor, 399–400, 406
natural fiber filled rubber transformation tensor, 351–352
composites, selected Ethylene-Propylene-Diene
published works, 378 Monomer rubber (EPDM)
optimum dispersion, 92–93 compounds, 38
properties of, 91 Ethylene-Propylene rubber (EPR), 45
reinforcing elastomers, 375–376 Evonik, see Degussa
resorcinol-HEXA system, 379 Experimental results
rubber-fiber bonding, 376 calculated data, comparison, 191–192
416 Index

Extraction kinetic method for bound Fillers


rubber assessment, 116 classification based on
Extrudate swell, see Postextrusion fabrication process and
swelling reinforcing activity, 12
particle sizes, 14
effect of
F
viscosity, 123–124, 333
Fabrication processes for carbon Yasuda parameter, 334
black (CB) yield stress, 333
aggregates, 21 fraction effect, 320
for furnace black, 22 incorporation effect, 125
lamp black, manufacturing loading and vulcanization,
process, 23 expected variation of
methods, 22 modulus function during, 126
smoke, 23–24 and pigment, distinguishing
thermal black process, 24 between, 13
thermo-oxidative processes, 21 refractive indices, 12
Face-centered cubic lattice model, 93 shapes, 11–12
FE, see Finite element (FE) calculations shear viscosity of, 95
Fiber-polymer systems, 341 structures, 12
Fiber-reinforced composites types, 11
shear-lag analysis, 389 use as color modifier, 12
Fiber-to-matrix modulus ratio, Filler–thermoplastic systems, 267
effect of, 391 Filler volume fraction
Fibrous fillers, 343 modulus, function, 220–221
Filled/composite polymer systems Finite element (FE) calculations, 353
classification, 7 Fitting data
Filled compounds with for filled polymer systems, 322–323
carbon black (CB) filled polystyrene (PS) shear
extraction kinetic data on N330, 117 viscosity, 326
morphology and nonlinear flow on PDMS/silica compound, 309–310
properties, 113 silica weight fraction
stress overshoot experiments on, 106 compound, 311–312
tridimensional representation virgin polystyrene (PS) data with
of morphology of, 111 Carreau–Yasuda model, 324
Filled polymers, 1 FKM, see Fluoroelastomers (FKM)
fitting experimental data, Flexural modulus, 404–405
322–324, 326 Floc simulation, 137
observations, 323 approach, 38–39
and polymer nanocomposites, 3 Flow anisotropy effects, 114
preparing and using, 2 in filled rubber compounds, 115
shear viscosity function, Flow properties, 263
model, 319–321 Flow zone, 125
filler fraction, effect of, 320 Fluoroelastomers (FKM), 52
Filled rubber compounds, flow Fluorohydric acid (HF), 56
anisotropy effects in, 115 Fractal geometry
Filler–filler network considerations, 171 solid volume fraction, 41
Filler network effect, 152; see also Fractal scaling law, 33
Dynamic stress softening (DSS) Fractional linear solid model, 222–223
Index 417

Fumed silica longitudinal (tensile)


combustion process, 58 modulus, 392
density, 58 random fiber orientation,
pyrolitic process, 56 modulus and adjustable
parameter, 394
transversal (tensile) and shear
G
modulus, 393
Gas black process, 23 Halpin–Tsai model, 361
Gaussian peaks, 47 commercial SGF-filled polyamide
Gaussian statistics, 251 6 and 66 composites, flexural
GCC, see Ground calcium and tensile moduli of, 362
carbonate (GCC) commercial SGF-filled polyamide
General purpose (GP) resins, 4 11 composites, flexural and
GenFit function, 324 tensile moduli of, 363
Glass fibers, 69 Hard clay, 257
types of, 71 Hard fibers, 72
Glass fibers reinforced polyesters, 1 Henry’s law, 46
Grain effect, 114 Herschel–Bulkley equation, 270
Green tire, 235 Herschel–Bulkley model for yield
Ground calcium carbonate (GCC), 55 stress fluids, 100–101
Ground talc, 260 HEXA, see Hexamethylenetetramine
Guth, Gold, and Simha equation, 122 (HEXA)
Guth and Gold approach, 138 Hexa(methoxymethyl)melamine
(HMMM), 377
Hexamethylenetetramine
H
(HEXA), 377
HAF, see High abrasion furnace (HAF) HF, see Fluorohydric acid (HF)
Halpin–Tsai equations High abrasion furnace (HAF), 28–29
curves calculations, 347 grade, 40
fiber aspect ratio, 350–351 High structure silica, 61
fiber orientation distribution, 348 HMMM, see Hexa(methoxymethyl)
filler particle’s geometry, ζ melamine (HMMM)
parameter expressions Hydrophilic fillers, 263
for, 346 Hydrophilic polymers, 4
fitting experimental data with, 348 Hydrous kaolin, 49
glass fibers, 348–349
mechanical properties, 346
I
Nielsen modification of Halpin, 396
longitudinal (tensile) IGC, see Inverse gas
modulus, 398 chromatography (IGC)
maximum packing functions, 397 Injection molded fatigue test samples
transverse (tensile) modulus and (ASTM D1708), 395
shear modulus, 398–399 in situ polimerization, 4
PBT and PA/PAT composites with Inter-aggregates distances
short glass fibers, 349 face-centered cubic lattice
short fibers filled systems, model, 93–94
347, 391–392 for optimal reinforcement, 94
average orientation Intermediate super abrasion furnace
parameters from, 394–396 carbon black (ISAF), 45
418 Index

International Organization for L


Standardization (ISO), 27
Ladouce–Stelandre model, 171
Intrinsic viscosity, 272, 335
Lamp black process, 23
Inverse gas chromatography (IGC), 46
Langmuir’s theory, 168
ISAF, see Intermediate super abrasion
Lattice gas model, 172
furnace carbon black (ISAF)
Layered silicates, 3, 5
ISO, see International Organization for
Lennard-Jones potential, 154
Standardization (ISO)
Liege, 12
Isolator–conductor transition, 172
Lignin, 75, 77
Lion model
J dynamic strain softening
(DSS), 223–225
Junction gap width, 189
Local drag flow mechanisms
in CB, 136
of rubber–aggregate
Junction rubber, 133
flow units, 114
Longitudinal (tensile) modulus, 392,
K 395, 398, 409
calculation, 402
Kaolin, 257
Low strain amplitude
grades, 245, 258–259
dynamic properties
Kaolins, see Clay(s)
silica reinforcement, 238
Kraus deagglomeration–
reagglomeration model
dynamic strain softening (DSS),
M
196–202, 209
complex modulus, 202 Magnesium hydroxide, 263
modeling, 197–200 Magnesium oxide
soft spheres interactions, 196 fluorohydric acid (HF), 56
Kraus model Maier and Göritz model, 170, 227
deagglomeration–reagglomeration development, 227–228
of filler aggregates, 155 experimental data, fitting, 230
G′ and G″ data, 157 mathematical aspects, 229
mathematical aspects, 204–205 modeling G′, strain, 231
rate equilibrium between modeling G″, strain, 232–233
dislocated and flocculated Mass fractal dimension of carbon
aggregates, 163 black (CB), 30, 33
SBR/carbon black compounds, 158 aggregate, volume of particle, 39
Ulmer modification, dynamic strain and aggregates, 33
softening (DSS), 212–215 connectivity exponent, 36
modeling G′, strain, 213–215 critical filler level, 43–44
Kraus–Ulmer equations, 253, 255 DPBA, 39–40
Krieger–Dougherty equation, ethylene-propylene-diene
272, 274–275 monomer rubber (EPDM)
polynomial, relative viscosity compounds, 38
variations, 276 fractal geometry, 36–37, 41
Krieger–Dougherty relationship fractal nature, 43
expansion, 335 geometry description of
numerical illustration, 336 aggregates, 34
Kuhn length, 246 HAF grade, 40
Index 419

Medalia, floc simulation Carreau–Yasuda model, 268–269, 278


approach, 38–39 Casson model, 101
Medalia aggregates, concept of, 37–38 for yield stress fluids, 102
Medalia occluded rubber, 42–43 cluster-cluster-aggregation (CCA)
pellets and agglomerates, 42 model, 158, 218–221
pellets mixing operations, 43 aggregate, solid fraction
TEM/AIA study, 37 of, 219–220
void ratio, 39 filler volume fraction,
volume fraction, 41–42 modulus, 220–221
well dispersed state, 43 strain dependence,
X-ray scattering experiments, 35 modulus, 221–222
Mastics, 247 Cohen–Addad percolation model
Matrix modulus, effect of, 390–391 maximum adsorbed polymer at
Maximum packing fraction, 315 saturation, 251
effect of, 317, 337 for silica-PDMS, 120
Medalia’s aggregate morphology comparison
approach, 86 sharp variation, 317
void ratio and DPB absorption, smooth variation, 317
relationship, 87–88 equivalent circuit model, 172
Medalia’s floc simulation, 193 face-centered cubic lattice
data, 85 model, 93–94
Mesophase concept, 150 fractional linear solid
Mesophase-pitch-precursor (MPP- model, 222–223
fibers), 71 Halpin–Tsai model, 361
Metamorphic rocks, 54 commercial SGF-filled polyamide
Methoxy group, 243 6 and 66 composites, flexural
Mica and tensile moduli of, 362
filled PP compounds, application, 298 commercial SGF-filled polyamide
and thermoplastics, 297 11 composites, flexural and
volume fraction, effect of, 300 tensile moduli of, 363
Michelin formulation, 246 Herschel–Bulkley model
Mineral fillers for yield stress fluids, 100–101
cost of, 2 Kraus deagglomeration–
industrial use, 264 reagglomeration model
Model equation, 326 dynamic strain softening
Modeling G′, strain, 213–215, 231–232 (DSS), 196–202, 209
Modeling G″, strain, 215–218, 232–233 Kraus model
Modeling, polymer systems deagglomeration–
shear viscosity function, 312–315 reagglomeration of filler
Modeling tan δ, γ0, 200–202 aggregates, 155
Models G′ and G″ data, 157
aggregates flocculation/ mathematical aspects, 204–205
entanglement model, 218–221 rate equilibrium between
aggregate, solid fraction dislocated and
of, 219–220 flocculated aggregates, 163
filler volume fraction, SBR/carbon black
modulus, 220–221 compounds, 158
strain dependence, Ulmer modification, dynamic
modulus, 221–222 strain softening (DSS), 212–215
420 Index

Ladouce–Stelandre model, 171 Mooney viscometer, 98


lattice gas model, 172 Mori–Tanaka’s average stress concept
Lion model Eshelby’s tensor, 399–400
dynamic strain softening experimental data, comparison,
(DSS), 223–225 404–405
Maier and Göritz model, 170, longitudinal (tensile) modulus
227–232 calculation, 402
development, 227–228 materials and volume fraction
experimental data, fitting, 230 depending constants, 401
mathematical aspects, 229 materials constants, 400–401
modeling G′, strain, 231 shear modulus calculation,
modeling G″, strain, 232–233 403–404
Mori–Tanaka’s averaging hypothesis transversal (tensile) modulus
and derived models calculation, 402–403
average strain approach, 352 Mori–Tanaka’s averaging hypothesis
fiber aspect ratio, 352–353 and derived models
finite element (FE) average strain approach, 352
calculations, 353 fiber aspect ratio, 352–353
liquid fillers, 353 finite element (FE)
Poisson’s ratio, 352 calculations, 353
thermoplastic polymer systems, liquid fillers, 353
flexural and tensile moduli Poisson’s ratio, 352
comparison, 353–354 thermoplastic polymer systems,
network junction (NJ) model, 134 flexural and tensile moduli
shear lag models, 353, 355 comparison, 353–354
efficiency factor, 355 Mullins effect, 127–128
fiber aspect ratio, 357
fiber lengths, distribution of, 358
N
packing ratio, 356
shear viscosity function Nanocomposites
model, 319–321 commercial demand, 5
Toom model, 48 fundamental research on, 6
Vand cubic model, 122 Hamed’s proposal, 6–7
van de Walle, Tricot and Nanometer-size materials, 4–5
Gerspacher (VTG) model, 165 Natural fiber filled rubber composites
White–Wang model, 101 selected published works, 378
for carbon black filled Natural fibers
compounds, 103 cellulose structure, 74, 77
CB filled rubber compounds, 102 composition, 72
Mohs hardness, 265 lignin, 75, 77
gypsum, 265 polymers
scale, 266 matrix interactions, chemical
soft minerals, 265 approaches, 78
wollastonite, 265 potential fillers for, 76
Molten polymer, 3 SEM microphotographs, 78
Monte Carlo method, 167 synthetic fibers, properties
Montmorillonite, 50 of, 72–73
Mooney and Rivlin equation, 130–131 Natural rubber (NR), 45, 382
Mooney functional analysis, 272 pseudo-Newtonian plateau, 95
Index 421

Natural rubber (RSS1) compounds, 237 silanization efficiency


Natural silica alkoxy groups, effect of, 243
neuburg silica, 52 3-thiocyanatopropyl-triethoxy
quartz, 52 silane (TCPTS), 240
siliceous fillers, 53 Orientation parameter, 396
diatomaceous earths, 53
white calcium silicate (CaSiO3),
P
52–53
Network junction (NJ) Packing fraction, 337
model, 134, 187 Packing of objects, 270
junction gap width, 189 Parallelepiped samples, 348
theory, strain amplification Particle/matrix adhesion, 284
factor, 190–194 Payne effect, 141
typical calculations, 188 PBT, see Polybutylene
strain amplification factor therephtalate (PBT)
experimental data, comparison, PCC, Precipitated Calcium
194–196 Carbonate (PCC)
Network junction (NJ) theory, 133 PDMS, see Polydimethylsiloxane
model development, 185–189 (PDMS)
absorbed DBP, 185 Percolation level, 46
NJ model, 185–187 Percolation theory, 153; see also
strain amplification factor and, 139 Dynamic stress
Neuburg silica, 52 softening (DSS)
Newtonian plateau, 267 Phyllosilicate, 3
Nielsen modification of Halpin, 396 Plan-plan rheometer, 96
longitudinal (tensile) modulus, 398 Plastics organoclay additives, 3
maximum packing functions, 397 Platy fillers, oriented, 260
transverse (tensile) modulus and Polyacrylonitrile (PAN-fibers), 71
shear modulus, 398–399 Polyamides, 4, 360
Nonlinear fitting algorithm, 309, 311 Polybutylene therephtalate (PBT), 348
NR, see Natural Rubber (NR); Polydimethylsiloxane (PDMS),
Natural rubber (NR) 237, 246–247
N330 SBR compound, 93 adsorption kinetics of, 250
Nuclear magnetic resonance chains, 248
(NMR), 248 chlorosilanes, hydrolysis of, 247
gel, 252
Polyethylene, 17–18
O
Polymers
OESBR/BR tread compound, 261 adsorption kinetics, model, 308
Organic ammonium salts, 4 cost of, 2
Organic fillers molecular weight, effect
chemical synthesis by, 12–13 of, 308–310
of natural origin, 12 PDMS/silica compound,
Organo-clays, 50–51 fitting data, 309
Organophilic clays, 4 natural fibers
Organophilicity, degree of, 237 matrix interactions, chemical
Organo-silanes, 236, 243, 261 approaches, 78
Bis(3-triethoxysilylpropyl) potential fillers for, 76
tetrasulfane (TESPT), 240 polymer–glass fiber systems, 150
422 Index

polymer-short fiber systems for R


industrial applications, 341
Reduced segmental mobility of rigid
and short fibers
body with interactions, 150
elastomers tensile properties
Reinforcement
difference, 339
fillers, 15
fiber-polymer systems, 341
structure of, 17
fibers-to-matrix adhesion
polyethylene or polypropylene, 18
role, 341
promoters, 240
fibrous fillers, 343
relative variation of rubber
harmonic mixing rule, 340–341
compound properties, 15–16
upper and lower bound
thermoplastics, 17–18
moduli curves, 340
Reinforcing elastomers, 375–376
Voigt average, 339–340
Relative viscosity of suspensions
silicone
comparing model equations, 273
polymer molecular weight,
Relaxation modulus function, 125
308–310
Rheology of suspensions of rigid
silica adsorption kinetics of,
particles, 315
308–312
Rigid particles suspensions
silica weight fraction, 310–312
rheology, models, 315–319
and short fibers
criteria, 315–316
filler, 341
Rigid PVC
Polynomial equation, 275
CaCO3 particle size effect, 290
Polyolefins, 2
Rigid spherical particles models, 125
Polyorganosiloxanes, 246
Rubber, 2
Polypropylene (PP), 287, 298
dynamic properties as tire
filler compounds, effect of
technology requirements, 140
talc volume fraction, 295–296
elastic behavior of
Polyvinyl chloride (PVC), 2, 18
modeling, 190–191
Postextrusion swelling, 106
matrix
and entrance pressure drop
CB aggregates, 186
in CB, 108
well dispersed state, 43
PP–CaCO3 composites, 281–282,
reinforcement, 24–25
286–287
silica silanisation reaction, type
mechanical properties, 284, 289
on, 244
Precipitated calcium carbonate (PCC), 55
technology, 91
Precipitated silica, 58
Rubber–carbon black interaction
amorphous, grades of, 58
chemical reticulation, 169
dispersion of, 61
hard and soft regions, 129
high structure silica, 61
stress–strain behavior, 131
manufacturing process of, 58–60
as topological constraints effect, 120
properties of, 60
Rubbery plateau, 125
synthetic, types of, 62
Probe, 46
PVC, see Polyvinyl chloride (PVC)
S
SAXS, see Small-angle X-ray
Q
scattering (SAXS)
Quartz SBR, see Styrene-butadiene
BET, specific surface area, 52 rubber (SBR)
Index 423

Scanning tunneling microscopy extrusion moving die


(STM), 35 technology, 380
Sealants, 247 fiber orientation, 382–383
Sedimentary rocks, 54 glass fibers exhibit, 381–382
SEM microphotographs, 78 hexa(methoxymethyl)melamine
SGF, see Short glass fibers (SGF) (HMMM), 377
Shape factor, 271 natural fiber filled rubber
Shear lag composites, selected
analysis, 389 published works, 378
models, 171, 409 natural origin
analysis, 344–345 commercial fibers-filled
efficiency factor, 355 polypropylene composites, 371
fiber aspect ratio, 357 wood flour and high stiffness
fiber lengths, distribution of, 371
of, 358 wood–polymer composites
longitudinal (tensile) (WPC), 371–374
modulus E11, 409 and polymers
packing ratio, 356 elastomers tensile properties
Shear modulus, 393, 398–399 difference, 339
calculation, 403–404, 408–409 fiber-polymer systems, 341
Shear viscosity fibers-to-matrix adhesion
complex cosmetic material, 98–99 role, 341
filled BR compound, 98 fibrous fillers, 343
function, effect on, 95 filler, 341
Herschel–Bulkley equation, 100 harmonic mixing rule, 340–341
plots and level, 97 upper and lower bound moduli
power law model with yield stress, curves, 340
99–100 Voigt average, 339–340
Shear viscosity function reinforcing elastomers, 375–376
capabilities, model resorcinol-HEXA system, 379
filled polymers, 319–321 rubber-fiber bonding, 376
critical shear rate and viscosity, 314 rubber matrix, 375–376
effect on carbon black (CB), 95 short fiber reinforced rubber
modeling, polymer systems, 312–315 composites, 379–380
mathematical aspects, 313 in extruded rubber hoses,
Short E-glass fiber controlling, 381
in different thermoplastics, special fiber–elastomer
effects of, 342 composites, 378–379
on heat distortion temperature styrene butadiene copolymer
of various thermoplastics, (SBR), 376–377
effects of, 343 technical constraints, 382
on impact resistance of various thermoplastic systems
thermoplastics, effects of, 343 micromechanic models for
Short fibers (synthetic) fiber, 382
carbon black filled SBR compound, Short fibers-filled polymer composites,
mechanical properties, 381 micromechanic models
cellulose fibers and chopped Halpin–Tsai equations, 345
aramid fibers industrial curves calculations, 347
importance, 380–381 fiber aspect ratio, 350–351
424 Index

fiber orientation distribution, 348 PA11, average suppliers’ data, 367


filler particle’s geometry, ξ PA66, average suppliers’ data, 366
parameter expressions for, 346 polyamide 6 and 66 composites,
fitting experimental data flexural and tensile
with, 348 moduli of, 362
glass fibers, 348–349 polyamide 11 composites,
mechanical properties, 346 flexural and tensile
PBT and PA/PAT composites moduli of, 362
with short glass fibers, 349 residual stresses, 363, 368
short fibers-filled systems, 347 commercial polypropylene,
minimum fiber length mechanical properties of, 360
homogeneous matrix, 344 E-glass fibers, 358, 360
load transfer, 344–345 filling polyamides with, 363–364
shear-lag analysis, 345 impact resistance, 361
Mori–Tanaka’s averaging hypothesis thermoplastic polymers
and derived models Bagley plots in capillary
average strain approach, 352 rheometry, 368
average stress/strain concept, 351 high fiber alignment, converging
Eshelby’s transformation flow results in, 369–370
tensor, 351–352 injection molding, 368–369
fiber aspect ratio, 352–353 slit extrusion of, 369
finite element (FE) Short synthetic fibers
calculations, 353 carbon fibers
liquid fillers, 353 aramid fibers, 71–72
moduli composite model, PAN-fibers and MPP-fibers, 71
explicit relationships, 351 glass fibers, 69
Poisson’s ratio, 352 types of, 71
shear lag models, 353, 355 Silanes
efficiency factor, 355 as coagents, benefits, 240
fiber aspect ratio, 357 Silanisation, 262
fiber lengths, distribution of, 358 efficiency, organo-silanes
packing ratio, 356 alkoxy groups, 243
Short fibers-filled systems, reinforcing properties, effect of
347, 391–392, 394–396 NR compounds, 244
average orientation parameters silica, 239
longitudinal (tensile) silica, reaction
modulus, 392, 395 rubber type, effect, 244
orientation parameter, 396 in situ, 241, 243, 245
transversal (tensile) ethanol formation, 243
modulus, 393, 396 problems, 241
random fiber orientation, modulus steps
and adjustable parameter, 394 silane condensation
Short glass fibers (SGF), 341 reaction, 242
aliphatic polyamides, filling, 360 silanol, alkoxy reaction, 242
commercial composites Silanols, 245
flexural and tensile moduli Silica
data for, 361 carbon black
PA6, average suppliers’ comparable series, 237
data, 364–365 tensile properties, 238
Index 425

carbon black, effect of talc


low strain dynamic lamellar surface, 51
properties, 239 Silicone polymers
fabrication processes silica, adsorption kinetics of,
fumed silica, 56, 58 308–312
precipitated silica, 58–62 polymer molecular weight,
modification, bi-functional 308–310
organosilane, 241–242 silica weight fraction, 310–312
network, 238 Siloxane group, 236
peculiar dynamic properties, 235 chains, adsorption
reinforcement empirical model, 248
low strain amplitude dynamic Simulation algorithms, 277
properties, 238 Small-angle x-ray scattering (SAXS), 35
silanisation, 239–240, 243, 246 Soft clays, 257
silanes, as coagents, 240 Soft spheres interactions, 196
silica/PDMS systems, 247–249 Spring-and-dashpot system, 164
adsorption, kinetics properties, 247 S-SBR, see Styrene butadiene
dynamic strain softening effect, copolymer (S-SBR)
modeling, 254–255 Stiffness (elasticity modulus), 292
low strain dynamic properties, STM, see Scanning tunneling
variation of, 256 microscopy (STM)
silica/polysiloxane system, 253 Stokes diameter, 93
surface chemistry of, 235–237, 249 Storage effects, 118
carbon black, comparison, 236 Storage modulus
surface energy aspects strain dependence, 221–222
chemistry of, 68–69 Strain amplification concept of
components for, 70 Mullins and Tobin, 132
dispersive and polar Strain amplification factor, 193–194
components, 69 NJ theory, 190–194
particles, 69 elastic behavior of a rubber,
weight fraction, effect of, 310–312 modeling, 190–191
fitting data, 311–312 Strain crystallization effect of NR, 127
Silicates Strain sweep experiments, 213, 230
calcined clays on SSBR compounds, 141
calcination steps, 51 Strain sweep tests, 178
clay Stress overshoot experiments
and chemical grafting of, 50–51 on carbon black filled compounds, 106
China clay, 49–50 stress overshoot effect, 105
hard and soft clays, Structural aspects and
distinction of, 49 characterization of CB
montmorillonite, 50 aggregates, 25–26
one-step and two-step grafting ASTM classification of, 29–30
technique, mechanism for, 51 Brunauer, Emmet, Teller (BET)
physical adsorption of method, 28
functional polymers, 50 cetyltrimethylammonium
polymer nanocomposites, 50 bromide CTAB adsorption
mica methods, 27–28
color, 52 dibutylphthalate (DBP) method, 28
muscovite, 51–52 DPA absorption method, 28
426 Index

elementary analysis, 26 coarse grades, 262


elementary particles and structure and elastomers
of aggregates, size, 28–29 fractured surfaces, 260
rubber reinforcement, 24–25 ground talc, 260
shapes of, 25 magnesium silicate, crystalline
standard characterization form, 291
methods, 27 minerals, 11
tread and carcass tire applications, 30 myriad products, 292
Styrene-butadience rubber (SBR), 45 platelets, basal surfaces, 261
Styrene butadiene copolymer (S-SBR) PP Composites, data, 292–294
anionic polymerization, properties, 292
production, 246 thermoplastics, application, 297
Styrene-butadiene rubber (SBR) volume fraction, effect of, 295–296
formulation, 91 Tandon–Weng expressions for
Surface energy aspects of randomly distributed
carbon black (CB) spherical particles
energetic sites, 47–48 Eshelby’s tensor, 406
esterification of, 48 materials and volume fraction
lamp and gas, chemical functions depending constants, 407
detection, 44 materials constants, 406–407
new equilibrium gas adsorption shear modulus, calculation,
techniques, 46–47 408–409
percolation level, 46 tensile modulus, calculation, 408
probe, 46 Tan δ variation, with strain
rubber–filler interactions, 45 amplitude and
rubber grade, 44 temperature, 142
surface activity, 44 TCPTS, see 3-Thiocyanatopropyl-
ToF-SIMS and XPS, 45 triethoxy silane (TCPTS)
Toom model, 48 TEM/AIA, see Transmission electron-
Surface properties, 263 microscopy/automated-
Swollen rubber–filler gel, 118 image-analysis (TEM/
Synthetic resins, 13 AIA) study
Synthetic silica, 53 Tensile measurements (ASTMD638),
ASTM classification, 64 395
characterization and structural Tensile modulus, 126
aspects of, 62–63 Tensile modulus calculation, 408
IGC techniques, 64 Tensile stress softening (TSS),
precipitated silicic acids (silica) 127–128
and active silicates, TESPT, see Bis(triethoxysilylpropyl)
properties of, 54 tetrasulfane (TESPT)
property ranges for, 63 Tetra organic phosphonium
silicates, 64 solutions, 4
suppliers data, 65–67 Theoretical model
usage in, 54 approximate
fitted equation,
comparison, 192
T
Thermal black process, 24
Talc, 260, 291 Thermo-activated reactions, 243
chemical modification, 261 Thermo-oxidative processes, 21
Index 427

Thermoplastics, 172 Tunnel effects, 173


effect of CB on Two-roll laboratory mill, 260
electrical conductivity, 175–177
rheological properties of, U
173–175
materials Unbound rubber, 112
polyamides, 360 Upper and lower bound moduli
processors, 358–359 curves, 340
natural origin
commercial fibers-filled V
polypropylene composites, 371
lignocellulosic fibers, polymer Vand cubic model, 122
composites preparation, 370 Van der Waals forces, 152
moisture diffusion, 370 van de Walle, Tricot and Gerspacher
wood flour and high stiffness (VTG) model, 165
of, 371 Vapor-grown carbon fibers
wood–polymer composites (VGCF), 176–177
(WPC), 371–374 Vegetal fibers, 12
thermoplastic polymer systems VGCF, see Vapor-grown carbon
flexural and tensile moduli, fibers (VGCF)
comparison, 353–354 Vicinal silanols, 236, 243
Thermoplastics and white fillers, 262 condensation, 236
and calcium carbonates, 280 Vinyl-based polymer, 3
and clay(s), 300–301 Viscous modulus, 166, 207, 212,
and mica, 297–300 224, 230
talc, 291 Void ratio, 39
application, 297 Medalia classification and
properties, 292 assumption, 43
3-Thiocyanatopropyl-triethoxy silane Volume Packing Fraction, 277
(TCPTS), 240 VTG, see van de Walle, Tricot and
Tightly BdR, 112 Gerspacher (VTG) model
Time-of-flight secondary ion mass Vulcanizable elastomers, 247
spectrometry (ToF-SIMS), 45 Vulcanization, 56
Tire technology requirements silica–rubber bonding, 241–242
for rubber dynamic system, 240
properties, 140 Vulcanized PDMS/silica systems
Titanium dioxide, 288 dynamic properties, 253
ToF-SIMS, see Time-of-flight
secondary ion mass
W
spectrometry (ToF-SIMS)
Toom model, 48 Western Europe consumption
Transmission electron-microscopy/ of rubbers, 2
automated-image-analysis White fillers
(TEM/AIA) study, 37 carbonates
Transversal (tensile) modulus, 393, 396 dolomite, 56
calculation, 402–403 GCC and PCC, 55
Transverse (tensile) modulus, 398–399 grades, PCC, 55
Tread tire applications, 30 concept of maximum
TSS, see Tensile stress softening (TSS) volume fraction, 266–268
428 Index

and elastomers, 235 Wood flour, 12


mineral fillers Wood–polymer composites
barium sulfate, 56 (WPC), 371
calcium oxide, 56–57 flexural properties, 373–374
zinc and magnesium oxide, 56 formulation, 374
natural silica injection molding and compression
neuburg silica, 52 molding, 372–373
quartz, 52 maleic anhydride-grafted, 375
siliceous fillers, 53 market amounts in North
white calcium silicate America, 11
(CaSiO3), 52–53 PP based composites, (pine)
silica, surface energy aspects wood flour effect of, 374
chemistry of, 68–69 production and processing,
components for, 70 problems, 373
dispersive and polar wood-filled thermoplastic
components, 69 composites, 372
particles, 69 WPC, see Wood-polymer
silica fabrication processes composites (WPC)
fumed silica, 56, 58
precipitated silica, 58–62
silicates X
calcined clays and talc, 51 XNBR, see Carboxylated nitrile
clays, 49–51 rubber (XNBR)
mica, 51–52 XPS, see X-ray photoelectron
synthetic silica, 53 spectroscopy (XPS)
characterization and structural X-ray photoelectron spectroscopy
aspects of, 62–67 (XPS), 45
precipitated silicic acids
(silica) and active silicates,
properties of, 54 Y
usage in, 54
thermoplastics properties, affect, 263 Yield stress data for filled rubber
White–Wang model, 101 compounds, 104
for carbon black filled
compounds, 103
Z
CB filled rubber compounds, 102
Wollastonite Zero-shear viscosity, 278
filled compounds, 53 Zinc oxide, 56
fluoroelastomers (FKM), 52 as weight predispersion in
grades, 52 EVA, 92

Das könnte Ihnen auch gefallen