You are on page 1of 684

MELT RHEOLOGY

AND ITS ROLE


IN PLASTICS
PROCESSING
THEORY AND APPLICATIONS

MELT RHEOLOGY
AND ITS ROLE
IN PLASTICS
PROCESSING
THEORY AND APPLICATIONS

JOHN M. DEALY

Department of Chemical Engineering


McGill University
Montreal, Canada
and

KURT F. WISSBRUN

Hoechst Celanese Research Division


Summit, New Jersey

ImiirI VAN NOSTRAND REINHOLD


~

_ _ _ _ _ _ New York

Copyright 1990 by Van Nostrand Reinhold


Library of Congress Catalog Card Number 89-29215
e-ISBN-13:978-1-4615-9738-4

ISBN-13 :978-1-4615-9740-7
DOl: 10.1007/978-1-4615-9738-4

All rights reserved. Certain portions of this work 1990 by


Van Nostrand Reinhold. No part of this work covered by the copyright
hereon may be reproduced or used in any form or by any means-graphic,
electronic, or mechanical, including photocopying, recording, taping,
or information storage and retrieval systems-without written permission
of the publisher.
Softcover reprint of the hardcover 1st edition 1990
Van Nostrand Reinhold
115 Fifth Avenue
New York, New York 10003
Van Nostrand Reinhold International Company Limited
11 New Fetter Lane
London EC4P 4EE, England
Van Nostrand Reinhold
480 La Trobe Street
Melbourne, Victoria 3000, Australia
Nelson Canada
1120 Birchmount Road
Scarborough, Ontario MIK 5G4, Canada
16 15

14 13

12 11

10 9 8 7 6 5 4 3

Library of Congress Cataloging-in-Publication Data

Dealy, John M.
Melt rheology and its role in plastics processing: theory and
applications/John M. Dealy and Kurt F. Wissbrun.
p.
cm.
Includes bibliographical references.
1. Plastics-Testing.
II. Title.
TA455.P5D28 1989
668.4'042-dc20

2. Rheology.

I. Wissbrun, Kurt F.
89-29215
CIP

Preface
This book is designed to fulfill a dual role. On the one hand it
provides a description of the rheological behavior of molten polymers. On the other, it presents the role of rheology in melt
processing operations. The account of rheology emphasises the
underlying principles and presents results, but not detailed derivations of equations. The processing operations are described qualitatively, and wherever possible the role of rheology is discussed
quantitatively. Little emphasis is given to non-rheological aspects of
processes, for example, the design of machinery.
The audience for which the book is intended is also dual in
nature. It includes scientists and engineers whose work in the
plastics industry requires some knowledge of aspects of rheology.
Examples are the polymer synthetic chemist who is concerned with
how a change in molecular weight will affect the melt viscosity and
the extrusion engineer who needs to know the effects of a change in
molecular weight distribution that might result from thermal degradation.
The audience also includes post-graduate students in polymer
science and engineering who wish to acquire a more extensive
background in rheology and perhaps become specialists in this area.
Especially for the latter audience, references are given to more
detailed accounts of specialized topics, such as constitutive relations
and process simulations. Thus, the book could serve as a textbook
for a graduate level course in polymer rheology, and it has been
used for this purpose.
The structure of the book is as follows. Chapter 1 is an introduction to rheology and to polymers for readers entering the field for
the first time. The reader is assumed to be familiar with the
mathematics and chemistry that are taught in undergraduate engineering and physical science programs.
Chapters 2 through 6 are a treatment of rheological behavior that
includes the well established areas of steady shear and linear
v

vi

PREFACE

viscoelasticity. There is, in addition, an extensive discussion of


nonlinear viscoelasticity effects, which often play an important role
in melt processing operations. Chapters 7 through 9 are devoted to
the experimental methods used to measure the properties that have
been defined, using both the traditional flows and some special
types of deformation.
The dependence of the parameters of the rheological relations
upon the composition and structure of the polymeric materials is
the subject of Chapters 10 through 13. The description is most
extensive for stable, homogeneous, isotropic molten polymers, and
less so for more complex systems. Chapters 14 through 17 summarize what is known about the role of rheology in the most important
melt processing operations. Finally, we close with a chapter whose
aim is to provide guidelines, often by example, of how to apply the
information in this book and in the literature to solve problems in
applied rheology.
This volume is not an exhaustive monograph on all aspects of
polymer rheology. However, we have included all the material that
we believe is likely to be of direct use to those working in the
plastics industry. The reference lists are not intended to be exhaustive, but all the work that we believe is central to the themes of the
book has been cited.
We have adhered to the Society of Rheology official nomenclature wherever possible. Also, we have used index rather than dyadic
notation for tensor quantities, because we felt this would be more
easily understood by readers seeing tensor notation for the first
time.
JMD wishes to acknowledge the support and encouragement of
McGill University for providing a working environment conducive
to a major writing project. He also wishes to recognize the colleagues and research students who have played a vital role in the
development of his understanding of polymer rheology and its
applications. In addition, JMD wishes to express his appreciation to
the University of Wisconsin, especially to R. B. Bird and A. S.
Lodge, for their professional hospitality during the time when he
got his part of the writing well launched.
KFW wishes to acknowledge the management of Hoechst
Celanese for their permission to participate in this book. He also

PREFACE

vii

wishes to thank his many colleagues at Hoechst Celanese, in particular H. M. Yoon, and his colleagues at the University of Delaware,
most especially A. B. Metzner, for their contributions to his experience and knowledge of the fields discussed in this book. Others to
whom appreciation is due include W. W. Graessley, F. N. Cogswell,
D. Pearson, M. Doi, and G. Fuller.
Several people read one or more chapters of the manuscript and
made many helpful suggestions for improvement. These include
H. M. Laun, 1. E. L. Roovers, H. C. Booij, G. A. Campbell, S. 1.
Kurtz, and 1. V. Lawler. Their contributions are gratefully acknowledged. Finally, we wish to thank Hanser Publishers, particularly Dr.
Edmund Immergut, for permission to reproduce some material
from our chapter in the Blow Molding Handbook.
J. M. Dealy
K. F. Wissbrun

Contents
Preface
1. INTRODUCTION TO RHEOLOGY

1.1
1.2
1.3
1.4

What is Rheology?
Why Rheological Properties are Important
Stress as a Measure of Force
Strain as a Measure of Deformation
1.4.1 Strain Measures for Simple Extension
1.4.2 Shear Strain
1.5 Rheological Phenomena
1.5.1 Elasticity; Hooke's Law
1.5.2 Viscosity
1.5.3 Viscoelasticity
1.5.4 Structural Time Dependency
1.5.5 Plasticity and Yield Stress
1.6 Why Polymeric Liquids are Non-Newtonian
1.6.1 Polymer Solutions
1.6.2 Molten Plastics
1.7 A Word About Tensors
1.7.1 Vectors
1.7.2 What is a Tensor?
1.8 The Stress Tensor
1.9 A Strain Tensor for Infinitesimal Deformations
1.10 The Newtonian Fluid
1.11 The Basic Equations of Fluid Mechanics
1.11.1 The Continuity Equation
1.11.2 Cauchy's Equation
1.11.3 The Navier-Stokes Equation
References

2. LINEAR VISCOELASTICITY

2.1 Introduction
2.2 The Relaxation Modulus

v
1
1
3
3
6
7
9
10
10
11
13
16
18
19
19
20
22
23
23
25
31
36
37

38
39
40
41
42

42
43
ix

CONTENTS

2.3 The Boltzmann Superposition Principle


2.4 Relaxation Modulus of Molten Polymers
2.5 Empirical Equations for the Relaxation Modulus
2.5.1 The Generalized Maxwell Model
2.5.2 Power Laws and an Exponential Function
2.6 The Relaxation Spectrum
2.7 Creep and Creep Recovery; The Compliance
2.8 Small Amplitude Oscillatory Shear
2.8.1 The Complex Modulus and the Complex
Viscosity
2.8.2 Complex Modulus of Typical Molten Polymers
2.8.3 Quantitative Relationships between G*(w) and
MWD
2.8.4 The Storage and Loss Compliances
2.9 Determination of Maxwell Model Parameters
2.10 Start-Up and Cessation of Steady Simple Shear and
Extension
2.11 Molecular Theories: Prediction of Linear Behavior
2.11.1 The Modified Rouse Model for Unentangled
Melts
2.11.1.1 The Rouse Model for Dilute Solutions
2.11.1.2 The Bueche Modification of the Rouse
Theory
2.11.1.3 The Bueche-Ferry Law
2.11.2 Molecular Theories for Entangled Melts
2.11.2.1 Evidence for the Existence of
Entanglements
2.11.2.2 The Nature of Entanglement Coupling
2.11.2.3 Reptation
2.11.2.4 The Doi-Edwards Theory
2.11.2.5 The Curtiss-Bird Model
2.11.2.6 Limitations of Reptation Models
2.12 Time-Temperature Superposition
2.13 Linear Behavior of Several Polymers
References

3. INTRODUCTION TO NONLINEAR VlSCOEIASTICITY


3.1
3.2

Introduction
Nonlinear Phenomena

44
48
51
52
53
54
55
60
61
66
68
69
70
72
74
74
74
75
79
79
79
80
81
82
85
86
86
94
100

103
103
105

CONTENTS

3.3 Theories of Nonlinear Behavior


3.4 Finite Measures of Strain
3.4.1 The Cauchy Tensor and the Finger Tensor
3.4.2 Strain Tensors
3.4.3 Reference Configurations
3.4.4 Scalar Invariants of the Finger Tensor
3.5 The Rubberlike Liquid
3.5.1 A Theory of Finite Linear Viscoelasticity
3.5.2 Lodge's Network Theory and the Convected
Maxwell Model
3.5.3 Behavior of the Rubberlike Liquid in Simple
Shear Flows
3.5.3.1 Rubberlike Liquid in Step Shear Strain
3.5.3.2 Rubberlike Liquid in Steady Simple
Shear
3.5.3.3 Rubberlike Liquid in Oscillatory Shear
3.5.3.4 Constrained Recoil of Rubberlike
Liquid
3.5.3.5 The Stress Ratio (N1/u) and the
Recoverable Shear
3.5.4 The Rubberlike Liquid in Simple Extension
3.5.5 Comments on the Rubberlike Liquid Model
3.6 The BKZ Equation
3.7 Wagner's Equation and the Damping Function
3.7.1 Strain Dependent Memory Function
3.7.2 Determination of the Damping Function
3.7.3 Separable Stress Relaxation Behavior
3.7.4 Damping Function Equations for Polymeric
Liquids
3.7.4.1 Damping Function for Shear Flows
3.7.4.2 Damping Function for Simple Extension
3.7.4.3 Universal Damping Functions
3.7.5 Interpretation of the Damping Function in Terms
of Entanglements
3.7.5.1 The Irreversibility Assumption
3.7.6 Comments on the Use of the Damping Function
3.8 Molecular Models for Nonlinear Viscoelasticity
3.8.1 The Doi-Edwards Constitutive Equation
3.9 Strong Flows; The Tendency to Stretch and Align
Molecules
References

xi

106
108
109
110
112
113
114
115
117
118
119
119
121
122
122
123
126
127
128
128
131
132
134
134
138
139
141
142
144
146
148
150
151

xii

CONTENTS

4. STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC


FUNCTIONS
4.1
4.2
4.3
4.4
4.5

Introduction
Steady Simple Shear Flow
Viscometric Flow
Wall Slip and Edge Effects
The Viscosity of Molten Polymers
4.5.1 Dependence of Viscosity on Shear Rate
4.5.2 Dependence of Viscosity on Temperature
4.6 The First Normal Stress Difference
4.7 Empirical Relationships Involving Viscometric
Functions
4.7.1 The Cox-Merz Rules
4.7.2 The Gleissle Mirror Relations
4.7.3 Other Relationships
References
5. TRANSIENT SHEAR FLOWS USED TO STUDY
NONLINEAR VISCOELASTICITY
5.1 Introduction
5.2 Step Shear Strain
5.2.1 Finite Rise Time
5.2.2 The Nonlinear Shear Stress Relaxation Modulus
5.2.3 Time-Temperature Superposition
5.2.4 Strain-Dependent Spectrum and Maxwell
Parameters
5.2.5 Normal Stress Differences for Single-Step Shear
Strain
5.2.6 Multistep Strain Tests
5.3 Flows Involving Steady Simple Shear
5.3.1 Start-Up Flow
5.3.2 Cessation of Steady Simple Shear
5.3.3 Interrupted Shear
5.3.4 Reduction in Shear Rate
5.4 Nonlinear Creep
5.4.1 Time-Temperature Superposition of Creep Data
5.5 Recoil and Recoverable Shear
5.5.1 Creep Recovery
5.5.1.1 Time-Temperature Superposition;
Creep Recovery

153
153
153
155
158
158
159
169
170
173
173
175
176
176

179
179
181
181
183
188
188
190
191
194
194
199
203
205
206
209
210
210
213

CONTENTS

5.5.2 Recoil During Start-Up Flow


5.5.3 Recoverable Shear Following Steady Simple
Shear
5.6 Superposed Deformations
5.6.1 Superposed Steady and Oscillatory Shear
5.6.2 Step Strain with Superposed Deformations
5.7 Large Amplitude Oscillatory Shear
5.8 Exponential Shear; A Strong Flow
5.9 Usefulness of Transient Shear Tests
References
6. EXTENSIONAL FLOW PROPERTIES AND THEIR
MEASUREMENT
6.1
6.2
6.3

Introduction
Extensional Flows
Simple Extension
6.3.1 Material Functions for Simple Extension
6.3.2 Experimental Methods
6.3.3 Experimental Observations for LDPE
6.3.4 Experimental Observations for Linear Polymers
6.4 Biaxial Extension
6.5 Planar Extension
6.6 Other Extensional Flows
References

7. ROTATIONAL AND SLIDING SURFACE RHEOMETERS


7.1 Introduction
7.2 Sources of Error for Drag Flow Rheometers
7.2.1 Instrument Compliance
7.2.2 Viscous Heating
7.2.3 End and Edge Effects
7.2.4 Shear Wave Propagation
7.3 Cone-Plate Flow Rheometers
7.3.1 Basic Equations for Cone-Plate Rheometers
7.3.2 Sources of Error for Cone-Plate Rheometers
7.3.3 Measurement of the First Normal Stress
Difference
7.4 Parallel Disk Rheometers
7.5 Eccentric Rotating Disks

xiii

214
215
217
218
219
219
225
228
228

231
231
232
237
238
241
249
258
260
263
265
266
269
269
270
270
274
275
275
277
278
279
281
283
284

xiv

CONTENTS

7.6 Concentric Cylinder Rheometers


7.7 Controlled Stress Rotational Rheometers
7.8 Torque Rheometers
7.9 Sliding Plate Rheometers
7.9.1 Basic Equations for Sliding Plate Rheometers
7.9.2 End and Edge Effects for Sliding Plate
Rheometers
7.9.3 Sliding Plate Melt Rheometers
7.9.4 The Shear Stress Transducer
7.10 Sliding Cylinder Rheometers
References
8. FLOW IN CAPILLARIES, SLITS AND DIES
8.1
8.2

8.3

8.4
8.5

8.6
8.7
8.8
8.9

Introduction
Flow in a Round Tube
8.2.1 Shear Stress Distribution
8.2.2 Shear Rate for a Newtonian Fluid
8.2.3 Shear Rate for a Power Law Fluid
8.2.4 The Rabinowitch Correction
8.2.5 The Schiimmer Approximation
8.2.6 Wall Slip in Capillary Flow
Flow in a Slit
8.3.1 Basic Equations for Shear Stress and Shear Rate
8.3.2 Use of a Slit Rheometer to Determine Nt
8.3.2.1 Determination of Nt from the Hole
Pressure
8.3.2.2 Determination of Nt from the Exit
Pressure
Pressure Drop in Irregular Cross Sections
Entrance Effects
8.5.1 Experimental Observations
8.5.2 Entrance Pressure Drop-the Bagley End
Correction
8.5.3 Rheological Significance of the Entrance
Pressure Drop
Capillary Rheometers
Flow in Converging Channels
8.7.1 The Lubrication Approximation
8.7.2 Industrial Die Design
Extrudate Swell
Extrudate Distortion

285
286
287
287
288
289
290
292
294
294
298
298
298
298
299
301
303
304
305
307
307
309
310
313
317
317
318
319
323
324
329
329
332
332
336

CONTENTS

8.9.1
8.9.2
8.9.3
8.9.4
8.9.5
References

Surface Melt Fracture-Sharkskin


Oscillatory Flow in Linear Polymers
Gross Melt Fracture
Role of Slip in Melt Fracture
Gross Melt Fracture Without Oscillations

9. RHEO-OPTICS AND MOLECUlAR ORIENTATION


9.1

Basic Concepts-Interaction of Light and Matter


9.1.1 Refractive Index and Polarization
9.1.2 Absorption and Scattering
9.1.3 Anisotropic Media; Birefringence and Dichroism
9.2 Measurement of Birefringence
9.3 Birefringence and Stress
9.3.1 Stress-Optical Relation
9.3.2 Application of Birefringence Measurements
References
10. EFFECTS OF MOLECUlAR STRUCTURE
10.1 Introduction and Qualitative Overview of Molecular
Theory
10.2 Molecular Weight Dependence of Zero Shear Viscosity
10.3 Compliance and First Normal Stress Difference
10.4 Shear Rate Dependence of Viscosity
10.5 Temperature and Pressure Dependence
10.5.1 Temperature Dependence of Viscosity
10.5.2 Pressure Dependence of Viscosity
10.6 Effects of Long Chain Branching
References
11. RHEOWGY OF MULTIPHASE SYSTEMS
11.1 Introduction
11.2 Effect of Rigid Fillers
11.2.1 Viscosity
11.2.2 Elasticity
11.3 Deformable Multiphase Systems (Blends, Block
Polymers)
11.3.1 Deformation of Disperse Phases and Relation to
Morphology

xv

337
338
339
340
341
341
345
345
346
347
349
353
358
358
362
363
365
365
368
370
374
381
381
384
386
389
390
390
390
392
400
401
403

xvi

CONTENTS

11.3.2 Rheology of Immiscible Polymer Blends


11.3.3 Phase-Separated Block and Graft Copolymers
References
12. CHEMORHEOLOGY OF REACTING SYSTEMS

12.1 Introduction
12.2 Nature of the Curing Reaction
12.3 Experimental Methods for Monitoring Curing Reactions
12.3.1 Dielectric Analysis
12.4 Viscosity of the Pre-gel Liquid
12.5 The Gel Point and Beyond
References
13. RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL
POLYMERS

13.1
13.2
13.3
13.4

Introduction
Rheology of Low Molecular Weight Liquid Crystals
Rheology of Aromatic Thermotropic Polyesters
Relation of Rheology to Processing of Liquid Crystal
Polymers
References
14. ROLE OF RHEOLOGY IN EXTRUSION

14.1 Introduction
14.1.1 Functions of Extruders
14.1.2 Types of Extruders
14.1.3 Screw Extruder Zones
14.2 Analysis of Single Screw Extruder Operation
14.2.1 Approximate Analysis of Melt Conveying Zone
14.2.2 Coupling Melt Conveying to Die Flow
14.2.3 Effects of Simplifying Approximations
14.2.3.1 Geometric Factors
14.2.3.2 Leakage Flow
14.2.3.3 Non-Newtonian Viscosity
14.2.3.4 Non-Isothermal Flow
14.2.4 Solids Conveying and Melting Zones
14.2.4.1 Feeding and Solids Conveying
14.2.4.2 Melting Zone
14.2.5 Scale-Up and Simulation

406
407
408
410

410
411
413
417
418
419
421

424

424
426
431
437
439
441

441
442
443
444
446
446
454
459
459
460
462
467
470
470
472
476

CONTENTS

14.2.5.1 Scale-Up
14.2.5.2 Simulation
14.3 Mixing, Devolatilization and Twin Screw Extruders
14.3.1 Mixing
14.3.2 Devolatilization
14.3.3 Twin Screw Extruders
References
15. ROLE OF RHEOLOGY IN INJECTION MOLDING
15.1
15.2
15.3
15.4

Introduction
Melt Flow in Runners and Gates
Flow in the Mold Cavity
Laboratory Evaluation of Molding Resins
15.4.1 Physical Property Measurement
15.4.2 Moldability Tests
15.5 Formulation and Selection of Molding Resins
References
16. ROLE OF RHEOLOGY IN BLOW MOLDING
16.1
16.2
16.3
16.4

Introduction
Flow in the Die
Parison Swell
Parison Sag
16.4.1 Pleating
16.5 Parison Inflation
16.6 Blow Molding of Engineering Resins
16.7 Stretch Blow Molding
16.8 Measurement of Resin Processability
16.8.1 Resin Selection Tests
16.8.2 Quality Control Tests
References
17. ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET
EXTRUSION
17.1 The Film Blowing Process
17.1.1 Description of the Process
17.1.2 Criteria for Successful Processing
17.1.3 Principal Problems Arising in Film Blowing
17.1.4 Resins Used for Blown Film

xvii

476
477
480
480
484
485
489
490
491
492
494
500
501
502
506
507
509
509
510
512
519
521
521
522
523
524
524
528
529

530
531
531
533
534
534

xviii

CONTENTS

17.2 Flow in the Extruder and Die; Extrudate Swell


17.3 Melt Flow in the Bubble
17.3.1 Forces Acting on the Bubble
17.3.1.1 Viscous Stress in the Molten Region of
the Bubble
17.3.1.2 Aerodynamic Forces
17.3.2 Bubble Shape
17.3.3 Drawability
17.4 Bubble Stability
17.5 Sheet Extrusion
References
18. ON-LINE MEASUREMENT OF RHEOLOGICAL
PROPERTIES
18.1 Introduction
18.2 Types of On-Line Rheometers for Melts
18.2.1 On-Line Capillary Rheometers for Melts
18.2.2 Rotational On-Line Rheometers for Melts
18.2.3 In-Line Melt Rheometers
18.3 Specific Applications of Process Rheometers
References
19. INDUSTRIAL USE OF RHEOMETERS
19.1 Factors Affecting Test and Instrument Selection
19.1.1 Purposes of Rheological Testing
19.1.2 Material Limitations on Test Selection
19.1.3 Instruments
19.2 Screening and Characterization
19.2.1 Advantages and Disadvantages of Rheological
Th~

19.2.2 Other Information Useful for Screening


19.2.3 Stability
19.2.3.1 Stability Measurement
19.2.3.2 Use of Stability Data
19.2.4 Temperature and Frequency Dependence
19.2.4.1 Measurement Tactics
19.2.4.2 Interpretation of Results
19.3 Resin Selection and Optimization and Process Problem
Solving

536
540
541

557
557
558
558
560
562
563
565
567
567
568
569
571
573

5n
574
577
578
580
582
582
583
585

CONTENTS

19.4 Rheological Quality Control Tests


References
APPENDIX A:

xix

595
599

MEASURES OF STRAIN FOR LARGE


DEFORMATIONS

601

MOLECULAR WEIGHT DISTRIBUTION AND


DETERMINATION OF MOLECULAR WEIGHT
AVERAGES

607

THE INTRINSIC VISCOSIlY AND THE


INHERENT VISCOSIlY

613

APPENDIX D:

THE GLASS TRANSITION TEMPERATURE

617

APPENDIX E:

MANUFACTURERS OF MELT RHEOMETERS


AND RELATED EQUIPMENT

622

APPENDIX B:

APPENDIX C:

NOMENCLATURE

630

AUTHOR INDEX

639

SUBJECT INDEX

649

MELT RHEOLOGY
AND ITS ROLE
IN PLASTICS
PROCESSING
THEORY AND APPLICATIONS

Chapter 1
Introduction to Rheology
1.1 WHAT IS RHEOLOGY?

It is anticipated that many readers will have little previous knowl-

edge about rheology but will wish to find out how it can be useful to
them in solving practical problems involving the flow of molten
plastics. For this reason, it is our intention to supply sufficient basic
information about rheology to enable the reader to understand and
make use of the methods described. With this in mind, we begin at
the beginning, with a definition of rheology.
Rheology is the science that deals with the way materials deform
when forces are applied to them. The term is most commonly
applied to the study of liquids and liquid-like materials such as
paint, catsup, oil well drilling mud, blood, polymer solutions and
molten plastics, i.e., materials that flow, although rheology also
includes the study of the deformation of solids such as occurs in
metal forming and the stretching of rubber.
The two key words in the above definition of rheology are
deformation and force. To learn anything about the rheological
properties of a material, we must either measure the deformation
resulting from a given force or measure the force required to
produce a given deformation. For example, let us say you wish to
evaluate the relative merits of several foam rubber pillows. Instinctively, you would squeeze (deform) the various products offered,
noting the force required to deform the samples.
A pillow that required a high force to compress would be considered "hard," and you probably wouldn't buy it, because it would be
painful to sleep on. On the other hand, if it required too little force
(too "soft") it would not provide adequate support for your weary
head. Foam rubber is a lightly crosslinked elastomer, and in squeez-

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

ing it you would be evaluating it primarily on the basis of its


resistance to deformation, i.e., its elastic modulus, which is a
rheological property.
Now let us say that you are looking for a lubricating oil for some
household application, and that your local hardware store has a
display in which samples of several oils are contained in glass
bottles. You can pick up the bottles and invert them to make the oil
flow from one end to the other. The "light" oils flow very rapidly
and splash as they hit the end of the bottle. The "heavy" oils creep
slowly in response to your test and take several seconds to accumulate at the end of the bottle. This is a rheological experiment. You
have used the earth's gravitational field to supply the force, and the
speed with which the oils flow is a measure of their rates of
deformation. Lubricating oils are viscous liquids, and in tipping the
bottles you would be evaluating them on the basis of their viscosity,
i.e., their resistance to flow, and this is another example of a
rheological property.
There are two principal aspects of rheology. One involves the
development of quantitative relationships between deformation and
force for a material of interest. The information for the development of such a relationship is obtained from experimental measurements. For example, for the foam pillow it might be observed that
the force required to compress it a certain distance is proportional
to the distance. In the case of the lubricating oil, it might be found
that the speed with which it flows through a small hole in the
bottom of a can is proportional to the height of the oil remaining in
the can.
For a linear elastic material or a Newtonian fluid, such simple
observations are sufficient to establish a general equation describing
how such a material will respond to any type of deformation. Such
an equation is called a "constitutive equation" or a "rheological
equation of state." However, for more complex materials such as
molten plastics, the development of a constitutive equation is a
much more difficult task, which may require the results of many
types of experiment.
The second aspect of rheology is the development of relationships that show how rheological behavior is influenced by the
structure and composition of the material and the temperature and
pressure. Ideally, one would like to know how these parameters
affect the constitutive equation, but this has not been accomplished

INTRODUCTION TO RHEOLOGY

at the present time, except for very simple materials such as


Newtonian fluids. In the case of more complex materials, one can at
least develop relationships showing how specific rheological properties such as the viscosity and the relaxation modulus are influenced
by molecular structure, composition, temperature and pressure.
Molten plastics are rheologically complex materials that can
exhibit both viscous flow and elastic recoil. A truly general constitutive equation has not been developed for these materials, and our
present knowledge of their rheological behavior is largely empirical.
This complicates the description and measurement of their rheological properties but makes polymer rheology an interesting and
challenging field of study.
1.2 WHY RHEOLOGICAL PROPERTIES ARE IMPORTANT

The forces that develop when a lubricant is subjected to a high-speed


shearing deformation are obviously of central importance to mechanical engineers. The rheological property of interest in this
application is the viscosity. The stiffness of a steel beam used to
construct a building is of great importance to civil engineers, and
the relevant property here is the modulus of elasticity.
The viscoelastic properties of molten polymers are of importance
to plastics engineers, because it is these properties that govern flow
behavior whenever plastics are processed in the molten state. For
example, in order to optimize the design of an extruder, the
viscosity must be known as a function of temperature and shear
rate. In injection molding, the same information is necessary in
order to design the mold in such a way that the melt will completely
fill it in every shot. In blow molding, the processes of parison sag
and swell are governed entirely by the rheological properties of the
melt.
1.3 STRESS AS A MEASURE OF FORCE

It was emphasized in Section 1.1 that force and deformation are the

key words in the definition of rheology. In order to describe the


rheological behavior of a material in a quantitative way, i.e., to
define rheological material constants (such as the viscosity of a
Newtonian fluid) or material functions (such as the relaxation
modulus of a rubber), it is necessary to establish clearly defined and

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

quantitative measures of force and deformation. Furthermore, it is


necessary to define these measures in such a way that they describe
the state of the material of interest without detailed reference to
the procedure used to make the rheological measurement.
For example, in the case of the evaluation of the pillows that was
described in Section 1.1, one way to quantify the test results would
be to place the pillow on a table, place a flat board on top, and
measure the distance between the board and the table both before
and after a weight of a certain mass was placed on top of the board,
as shown in Figure 1-1. If our objective is simply to compare several
pillows of the same size, it would be sufficient to simply list the
amount of compression, in centimeters, caused by a weight having a
mass of 1 kg.

Figure 1-1. Setup for testing pillows.

However, if our objective is to make a quantitative determination


of the elastic properties of the foam rubber, the reporting of the
test results is awkward. We must report the size and shape of the
sample (the pillow), the mass of the weight applied, and the amount
of compression. If one wishes to compare the behavior of this foam
with that of a second foam, the second material must be tested in
exactly the same way as the first. It would be advantageous to be
able to describe the elastic behavior of the rubber using physical
quantities which are defined so that they describe the state of the
material under test, without reference to the details of the test
procedure.
First let's look at force. Two types of force can act on a fluid
element. A "body force" acts directly on the mass of the element as
the result of a force field. Usually only gravity need be considered,
but a magnetic field can also generate a body force. A surface force
is the result of the contact of a fluid element with a solid wall or
with the surrounding fluid elements. It is the surface forces that are
of interest in rheology. The force exerted by a weight sitting on top

INTRODUCTION TO RHEOLOGY

Figure 1-2. Uniaxial (simple) extension.

of a pillow is an example of a surface force. The fact that the


ultimate cause of the surface force acting on the pillow is the body
force acting on the weight is not of rheological importance, as the
compressive force on the sample could equally well be supplied by
means of a testing machine, and the observed relationship between
force and deformation would be the same.
Placing a weight having a mass of 1 kg on a small pillow will
cause more compression than placing it on a larger pillow. From
the point of view of the material, it is obviously not the total force
that is important. In fact, since the deforming force acts on the
upper surface of the sample, it is found that if the force is divided
by the area of the surface we obtain a quantity suitable for describing the properties of the material. We call this quantity the "stress."
In general, then, the stress is calculated by dividing the force by
the area over which it acts. In the case of a test like the one with
the pillow which involves squeezing, we call this the compressive
stress. A more common type of test method for elastic materials
involves stretching rather than compressing, as shown in Figure 1-2.
Again, the stress is the force divided by the cross sectional area of
the sample, and in this case we call it a "tensile stress." We will use
the symbol (FE for this quantity.

(FE

stretching force
cross sectional area

(1-1)

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

is,,-

--l

f:j.X \ -

; - WETTED AREA

~A
F

--.

Figure 1-3. Simple shear (2 plates, gap

~ h).

Compressive and tensile stresses are the two types of "normal


stress," so called because the direction of the force is normal
(perpendicular) to the surface on which it acts. In addition to
normal stresses, we can have a "shear stress"; in this case, the
direction of the force is tangential to the surface on which it acts, as
shown in Figure 1-3. This figure shows the deformation called
"simple shear," in which the sample is contained between two fiat
plates with a fixed spacing, h, between them. The upper plate
moves in a direction parallel to itself while the lower plate is
stationary. The shear stress is the shear force divided by the
tangential area. We will use the symbol (T, with no subscript, to
refer to the shear stress in a simple shear deformation.
shear force
(T=

------------

tangential area

(1-2)

1.4 STRAIN AS A MEASURE OF DEFORMATION

In the previous section, we stated that shear stresses and normal


stresses are useful measures of the forces that act to deform a
material. Now we need a quantitative measure of deformation that
is rheologically significant. The description of deformation in terms
of strain is more complex than the description of force in terms of
stress, and there are many alternative, rheologically significant,
measures of strain. While we will consider this question further in
Chapter 3, we will define here only those measures that are useful

INTRODUCTION TO RHEOLOGY

in the description of deformations commonly used to make rheological measurements, namely simple shear and simple extension. In
Section 1.9 a strain measure for small deformations that is not
restricted to describing simple shear or simple extension will be
defined.
The thing that complicates the definition of a measure of strain is
that it is necessary to refer to two states of a material element. In
other words, it is not possible to specify the strain of a material
element unless we specify at the same time the reference state
relative to which the strain is measured. In the case of an elastic
material that cannot flow, for example a crosslinked rubber, this is
straightforward, because there is a unique, easily identifiable, unstrained state that a material element will always return to whenever deforming stresses are not acting.
For materials that flow, i.e., fluids, such a unique reference state
does not exist. In the case of a well-controlled experiment, however,
in which a simple homogeneous deformation is imposed on a
sample initially at rest and free of all deforming stresses, this initial
condition provides a meaningful reference state with respect to
which strain can be defined. We will make use of this fact in the
next two sections to define strain measures for simple extension and
simple shear.
1.4.1 Strain Measures for Simple Extension

Consider the simple extension test illustrated in Figure 1-4. Let Lo


be the length of the sample prior to the application of a tensile
stress and L the length after deformation has occurred. A simple
measure of the deformation is the quantity (L - Lo). However, this
quantity is meaningful only in terms of a specific sample, whereas
we desire a measure of deformation that describes the state of a
material element. We can easily form such a quantity by dividing
this length difference by the initial length to obtain the "linear
strain" for simple extension.

(l-3)
For a uniform deformation, every material element of the sample
experiences this same strain. For example, if the initial length at
time to of a material element measured in the direction of stretch-

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

8X,(t)

Figure 1-4. Quantities used to describe simple extension.

ing is oX1(t O)' and the length at a later time, t, after deformation
has occurred is OX/f), the linear strain of the material element is:

(1-4)
This measure of deformation has some convenient features. It is
independent of sample size, and it is zero in the unstressed, initial
state.
However, it is not the only measure of deformation that has these
properties. Another is the Hencky strain, which is defined as
follows in terms of the length of a material element.

(1-5)
For a sample with initial length Lo undergoing uniform strain the
Hencky strain can also be expressed as:

(1-6)
For materials that flow, e.g., molten plastics, this quantity is more
useful than the linear strain. In fact, the linear and Hencky strains

INTRODUCTION TO RHEOLOGY

become equivalent in the limit of very small strains. This can be


demonstrated for simple extension by noting that e = In{1 + S) and
that the first term of the series expansion of In{1 + S) is S.
The Hencky strain rate is also a useful quantity for describing
rheological phenomena in simple extension. This is defined in
Equation 1-7 in terms of the length, L, of the sample.

i:

d In(L)/dt

(1-7)

We note that the initial length does not enter into the Hencky
strain rate but does enter into the linear strain rate dS / dt.
1.4.2 Shear Strain

Now consider simple shear, which is the type of deformation most


often used to make rheological measurements on fluids. Referring
to Figure 1-3, an obvious choice of a strain measure is the displacement of the moving plate, ax, divided by the distance between the
plates, h.
'Y =

aX/h

(1-8)

Referring to the two material particles shown in Figure 1-5 rather


than to the entire sample, we can define the shear strain, 'Y, for the
fluid element located at (Xl' X 2 , X 3 ) as

(1-9)
where aXI is the distance, measured in the Xl direction, between
two neighboring material particles that are separated by a distance

Figure 1-5. Two material particles in simple shear.

10

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

in the x 2 direction. In the absence of edge effects, i.e., for a


uniform deformation, every fluid element will undergo the same
strain, and the local shear strain will everywhere be equal to the
overall sample strain:
~X2

(1-10)
And the shear rate is simply the rate of change of the shear strain
with time:
1 dX

Y=hdt=-;;

(1-11)

where V is the velocity of the moving plate.


1.5 RHEOLOGICAL PHENOMENA

In this section we will examine the general types of rheological


behavior that can be exhibited by materials. These are elasticity,
viscosity, viscoelasticity, structural time dependency, and plasticity.
Although we will use simple extension and simple shear behavior in
this section to describe rheological phenomena, it is important to
remember that for rheologically complex materials such as polymeric liquids, the behavior observed in these simple deformations
does not tell the whole rheological story.
1.5.1 Elasticity: Hooke's Law

Elasticity is a type of behavior in which a deformed material returns


to its original shape whenever a deforming stress is removed. This
implies that a deforming stress is necessary to produce and maintain any deviation in shape from the original (unstressed) shape,
i.e., to produce strain. The simplest type of elastic behavior is that
in which the stress required to produce a given amount of deformation is directly proportional to the strain associated with that
deformation. For example, in simple extension this can be expressed as:

(1-12)

INTRODUCTION TO RHEOLOGY

11

The constant of proportionality, E, is called Young's modulus. This


relationship is called Hooke's law. The corresponding form of
Hooke's law for simple shear is:
er

Gy

(1-l3)

where G is the shear modulus or modulus of rigidity. We note that


in a purely elastic material like this, all the work done to deform
the material is stored as elastic energy and can be recovered when
the material is permitted to return to its equilibrium configuration.
Another way of describing elastic behavior is to specify the strain
that results from the application of a specific stress. For a Hookean
material we have, for simple shear:
y = Jer

(1-14)

where J is the shear compliance. Obviously, for a material following


Hooke's law:
J

I/G

( 1-15)

1.5.2 Viscosity

Viscosity is a property of a material that involves resistance to


continuous deformation. Unlike elasticity, the stress is not related
to the amount of deformation but to the rate of deformation. Thus
it is a property peculiar to materials that flow rather than to solid
materials. We will consider first the simplest type of rheological
behavior for a material that can flow. For simple shear this type of
behavior is described by a linear relationship between the shear
stress and the shear rate:
er = 'Y1Y

(1-16)

where 'Y1 is the viscosity. A material that behaves in this way is


called a Newtonian fluid.
For a Newtonian fluid, the viscosity is a "material constant,"
in that it does not depend on the rate or amount of strain. Single
phase liquids containing only low molecular weight compounds are

12

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1)0

>r-

OO

oo

(f)

'>

SHEAR RATE (1)

Figure 1-6. Viscosity-shear rate curve for a typical molten polymer.

Newtonian for all practical purposes. For multiphase systems, for


example suspensions and emulsions, and for polymeric liquids, the
relationship between stress and strain rate is no longer linear and
cannot be described in terms of a single constant. It is still convenient, however, to present the results of a steady simple shear
experiment in terms of a viscosity function 1]( y) defined as follows:
(1-17)

where a is the shear stress and y is the shear rate. A typical


viscosity-shear rate curve for a molten polymer is shown in Figure
1-6. The important features of this curve are listed below.
1. At sufficiently low shear rates, the viscosity approaches a
limiting constant value 1]0 called the zero shear viscosity.
2. The viscosity decreases monotonically as the shear rate is
increased. This type of behavior is called shear thinning. (An
older terminology is "pseudoplastic.")
3. At sufficiently high shear rates the viscosity might be expected
to level off again, although a high-shear-rate limiting value is
not observed in melts, because viscous heating and polymer
degradation usually make it impossible to carry out experiments at sufficiently high shear rates. Specific forms for the
viscosity function are presented in Chapter 4.

INTRODUCTION TO RHEOLOGY

13

1.5.3 Viscoelasticity

Polymeric materials, including solutions, melts, and crosslinked


elastomers, exhibit both viscous resistance to deformation and
elasticity. In the case of a vulcanized (crosslinked) rubber, flow is
not possible, and the material has a unique configuration that it will
return to in the absence of deforming stresses. However, the viscous
resistance to deformation makes itself felt by delaying the response
of the rubber to a change in stress. To illustrate this point, consider
the phenomenological analog of a viscoelastic rubber shown in
Figure 1-7. This mechanical assembly consists of a spring in parallel
with a dashpot. The force in the spring is assumed to be proportional to its elongation, and the force in the dashpot is assumed to
be proportional to its rate of elongation. Thus, the spring is a linear
elastic element, in which the force is proportional to the extension,
X, and the dashpot is a linear viscous element in which the force is
proportional to the rate of change of X. Note that this assembly
will always return to a unique length, the rest length of the spring,
when no force is acting on it. This assembly, called a Voigt body, is
not intended to be a physical or quantitative model for a rubber.

Figure 1-7. Voigt body analog of a viscoelastic solid.

14

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

However, the qualitative characteristics of its response to changes


in force are similar in some ways to those exhibited by rubbers.
Consider how this assembly would respond to the sudden application of a tensile load, F. This is called a "creep test". The force,
F, is the sum of the force in the spring, KeX, and that in the
dashpot, Kv(dX/dt). Thus:
(1-18)
We note that some of the work put into the assembly to deform it is
dissipated in the dashpot, while the remainder is stored elastically
in the spring. If X is initially zero, and the force F is suddenly
applied at time t = 0, this differential equation can be solved to
yield:
(1-19)
The important point to note is that the viscous resistance to
elongation introduces a time dependency into the response of the
assembly, and that this time dependency is governed by the ratio
(KjK e), which has units of time. If we take the force to be
analogous to the deforming stress in a viscoelastic material, and the
elongation to be analogous to strain, we see that a viscoelastic
rubber has a time constant and cannot respond instantaneously to
changes in stress. This is called a "retarded" elastic response. As
the time constant approaches zero, the behavior becomes purely
elastic.
Now we turn to the case of an elastic liquid. To illustrate certain
qualitative features of the rheological behavior of such a material,
consider the mechanical assembly shown in Figure 1-8. This assembly, called a Maxwell element, consists of a linear spring in series
with a linear dashpot. Note first that unlike the Voigt body, this
assembly has no unique reference length and will deform indefinitely under the influence of an applied force, assuming the dashpot is infinite in length. This is analogous to the behavior of an
uncrosslinked polymeric material above its glass transition and
melting temperatures. Such a material will flow indefinitely when
subjected to deforming stresses.
Now we examine the force on the Maxwell element when it is
subjected to a sudden stretching by an amount XO' The force, but

INTRODUCTION TO RHEOLOGY

15

Figure 1-8. Maxwell element analog of a viscoelastic liquid.

not the displacement, is the same in both the spring and dash pot.
Thus:
(1-20)

Again we note that some of the work done is dissipated in the


dashpot and the remainder is stored in the spring. The total
displacement of the assembly, Xo, is the sum of Xe and Xv:
(1-21)

Thus:
(1-22)

This ordinary differential equation can be solved to yield


(1-23)

Note that (Kv/Ke) is a time constant. The force thus decays or


relaxes exponentially. If we take the force to be analogous to the
deforming stress in an elastic liquid and the elongation to be

16

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

analogous to the strain, this process is analogous to a stress relaxation experiment. As in the case of the viscoelastic rubber, we note
that the combination of viscous and elastic properties endows the
material with a characteristic time and makes its response timedependent. As this "relaxation time" becomes shorter and shorter,
however, it becomes more and more difficult to devise an experiment that will reveal the elastic nature of the liquid, and its
behavior appears more and more like that of a purely viscous
material.
When we examine the rheological behavior of actual polymeric
materials, we find creep and relaxation behavior that is qualitatively
like those described above. In particular, the response to a sudden
change in stress or strain is always time dependent, never instantaneous, and there is both elastic storage of energy and viscous
dissipation. On the other hand, the creep and relaxation curves
cannot be described by a single exponential function involving a
single characteristic time. 1 As is explained in Section 2.5, however,
practical use can still be made of the concept of a relaxation time
by describing the viscoelastic behavior of real materials in terms of
a spectrum of relaxation times.
1.5.4 Structural Time Dependency

In our discussion of the viscosity function, we took the shear stress


to be independent of time at constant shear rate. For a Newtonian
fluid this is appropriate, because the stress responds instantaneously to the imposition of a constant shear rate. However, nonNewtonian fluids may not respond instantaneously so that when the
shearing deformation is begun, there is a transient period during
which the shear stress varies with time, starting from zero and
finally reaching a steady state value that can be used to calculate
the viscosity by use of Equation 1-17. The origin of this time
dependency may be a flow-induced change in the structure of the
fluid, as in the case of a concentrated suspension of solid particles.
For example, the state of aggregation of the suspended particles
can be changed significantly by shearing. This "structural time
IOther deficiencies of these simple analogs are that the Voigt body does not exhibit stress
relaxation and the Maxwell element does not exhibit retarded creep.

INTRODUCTION TO RHEOLOGY

17

m
w

II:
I-

m
II:
<{

SHEARING STARTED

TIME

Figure 1-9. Shear stress versus time for a material with structural time dependency.

dependency" is entirely dissipative, that is to say there is no elastic


energy storage, and all the work done to deform the material is
converted immediately into "heat," or more precisely, into internal
energy. Time dependency may also arise from changes in the
conformation and orientation of molecules, as in the case of polymer solutions, but this type of time dependency is a viscoelastic
effect rather than a structural time dependency.
Restricting attention here to inelastic materials such as suspensions of rigid particles in Newtonian fluids, in which no elastic
energy storage is possible, the most common type of structural time
dependency is that shown in Figure 1-9, where the shear stress
decreases with time after the start up of steady shear flow. This
type of behavior, in which the viscosity decreases with time, is
called "thixotropy." If the shearing is stopped, the viscosity will
gradually increase as the particles respond to inter-particle forces in
the absence of a shear stress, but the time required for the complete reformation of the equilibrium structure can be very long.
There are some materials, such as corn starch suspensions, that
exhibit the opposite effect, viz., the stress increases with time after
the start up of steady shear. This type of behavior is called
"rheopexy" or "antithixotropy." In the case of thixotropy, the shear
tends to break down structure that has been established in the
unsheared material, while in the case of a rheopectic material, the
shearing promotes the buildup of structure.

18

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1.5.5 Plasticity and Yield Stress

The structure in a concentrated suspension can be sufficiently rigid


that it permits the material to withstand a certain level of deforming stress without flowing. The maximum stress that can be sustained without flow is called the "yield stress," and this type of
behavior is called "plasticity." Metals generally exhibit plasticity, as
do semicrystalline polymers at temperatures between their melting
and glass transition temperatures. Highly filled melts are also
thought to have a yield stress, although precise measurement of this
property is difficult.
The simplest type of plastic behavior is that in which the excess
stress, above the yield stress, is proportional to the shear rate. For
simple shear flow, this type of behavior is described by Equation
1-24.
(1-24)

Here, (To is the yield stress and 7Jp is the plastic viscosity. A material
that behaves in this way is called a "Bingham plastic." This is an
idealized type of behavior that is not precisely followed by any real
material, but it is sometimes a useful approximation to real behav-

,.....
~

en
en
w
a:
Ien
a:
<C
w
I
en

SHEAR RATE (-()

Figure 1-10. Shear stress versus shear rate for a Bingham plastic, a Newtonian fluid, and a
shear thinning fluid.

INTRODUCTION TO RHEOLOGY

19

ior. Note that Equation 1-24 is not a constitutive equation, as it


does not describe all the components of the stress tensor in any
type of deformation but only the shear stress in simple shear.
Figure 1-10 compares shear stress versus shear rate curves for a
Bingham plastic, a Newtonian fluid and a shear thinning fluid.
1.6 WHY POLYMERIC LIQUIDS ARE NON-NEWTONIAN

It is important in applied polymer science to be able to relate


physical properties, including rheological properties, to molecular
structure. This subject is taken up in some detail in Chapters 2, 4
and 10. We will mention here only the general mechanisms by
which polymeric molecules endow liquids with complex rheological
behavior.
First, we should examine the question of why polymer molecules
are elastic. Our physical picture of a polymer molecule is that of a
long chain with many joints allowing relative rotation of adjacent
links. The presence of this large number of joints makes the
molecule quite flexible and allows many different configurations of
the molecule. At temperatures above the glass transition temperature a molecule will continually change its configuration due to
Brownian motion, but we can describe the state of a large number
of molecules in terms of statistical averages. For example, at a given
temperature there will be a unique average value of the end-to-end
distance, R, for the molecules of a polymeric liquid that has been at
rest for a sufficient length of time that it is in its equilibrium state.
Deforming the liquid will alter this average length, but if the
deformation is stopped, Brownian motion will tend to return the
average value of R to its equilibrium value. This is the molecular
origin of the elastic and relaxation phenomena that occur in polymeric liquids.
1 .6.1 Polymer Solutions

We consider first the behavior of a dilute solution in which the


forces acting on the polymer molecule are primarily those due to
the flow of the solvent. This situation is much simpler than that
existing in concentrated solutions and melts, where the rheological
behavior is governed by interactions between polymer molecules. In

20

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the absence of a velocity gradient in the solvent, the polymer


molecules will have an average end-to-end distance that is small,
and a distribution of orientations that is random. When the solution
undergoes deformation, a dissolved molecule will be dragged along
with it, but the polymer molecule is so large that different parts of it
will be exposed to different solvent velocities. In other words, the
solvent moves relative to the segments of the polymer molecule and
exerts forces on these segments that change the shape (conformation) of the molecule, increasing the average end-to-end distance
and promoting some degree of orientation in a preferred direction.
For example, in simple extension, the preferred direction is the
stretch direction.
Some types of deformation have a greater ability to increase R
and orient molecules than others. Deformations that can generate a
high degree of stretching and orientation are said to be "strong
flows," while those that cannot are said to be "weak."
We have seen how the deformation of a polymeric liquid can
alter the configuration and orientation of the molecules. Thus, the
physical nature of the fluid is altered by the deformation, and this,
in turn, alters the mechanical (rheological) properties of the fluid.
This explains, for example, why the viscosity of a polymer solution
depends to some extent on shear rate. At high shear rates, the
shape of the polymer molecules is different from that at low shear
rates, and this alters the solution's resistance to flow and thus its
viscosity. Furthermore, the partial orientation of the molecules, i.e.,
the shift of the average orientation of the R vectors away from
zero, introduces a strain-induced anisotropy that is responsible for
differences in the normal stresses in steady simple shear. Finally,
the tendency of Brownian motion to return the system of molecules
to its most probable, i.e., equilibrium, configuration distribution
explains why these fluids are elastic and have relaxation times.
1.6.2 Molten Plastics

In lhe case of concentrated polymeric liquids, including melts, the


transmission of deforming stresses to a molecule is primarily (in the
case of a concentrated solution) or totally (in the case of a melt)
due to interactions between polymer molecules rather than interactions between the solvent and the molecules. The detailed nature of
these interactions is not completely understood at this time, but it is

INTRODUCTION TO RHEOLOGY

21

observed that the short term response to rapid deformations of high


molecular weight molten polymers is very similar to that of a
crosslinked rubber. This has inspired the concept of a "temporary
network" that exists in the melt and acts like a rubbery network at
shorter times but whose "junctions" can slip over longer periods of
time to permit flow. The network is sometimes said to arise from
"entanglements" in the melt. However, the modern view is that the
rubbery behavior of melts is due not to an actual looping or
knotting of molecules around each other but simply to the constraints on their motion resulting from the fact that molecules
cannot cut through each other.
Entanglements occur because of the high degree of spatial overlap of the molecules. The existence of overlap is readily demonstrated by considering the measured size of the polymer coils. One
measure of molecular size is the "radius of gyration," R g For
linear polyethylene, Rg depends on the molecular weight, M as
follows:
Rg(cm) = 4 X 10- 9 X M1/2
For a polyethylene with a molecular weight of 10 6 glmol, the
volume of the sphere occupied by one molecule is therefore about
2.6 X 10- 16 cm3 The mass of this coil is 10 6 divided by Avogadro's
number, or 1.7 X 10- 18 g. The density of the coil in its occupied
volume is thus less than 0.01 g/cm 3 The observed melt density of
about 0.77 g/cm 3 can only be accounted for if parts of many other
coils are present in the volume occupied by this coil.
A similar calculation for a polymer with a molecular weight of
10 4 glmol gives a density of 0.1 g/cm 3 , which is considerably closer
to the measured bulk density. This shows that the degree of coil
overlap, and therefore the entanglement density increases sharply
with molecular weight.
Rubbery behavior occurs in a melt when the molecular weight is
above some critical value that varies from one polymer to another.
Above this molecular weight the number of entanglements becomes
sufficient to produce strong rubberlike effects.
The macroscopic effects of the strong interactions between polymer molecules in a melt include high viscosity and high elastic
recoil, especially just above the melting point or, in the case of an
amorphous polymer, just above the glass transition temperature. At
the same time, the nature of this strong interaction can be altered

22

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

temporarily by deformation so that high molecular weight melts


have highly nonlinear properties; for example, the viscosity is a very
strong function of the shear rate.
Polymeric materials are said to have a "memory," in that when
deforming stresses are eliminated, they tend to return to a previous
configuration. Crosslinked polymers have a "perfect memory" in
the sense that since their network is based on permanent chemical
crosslinks they always return to a unique equilibrium configuration,
whereas molten polymers are said to have a "fading memory," since
the entanglement network is not permanent and is altered by flow
and relaxation processes.

1.7 A WORD ABOUT TENSORS

For those readers who have had little if any experience in the use of
tensor notation, the very word "tensor" probably suggests a mathematical system of impenetrable mystery. However, such readers
should have no fear. There is no mystery! While we do not claim to
offer here a complete course in tensor analysis, we do present in the
next two brief sections everything you will need to know about
tensors in order to describe the rheological properties of polymeric
liquids. After a careful reading of these sections, you too can
impress the uninitiated with your ability to use tensor notation to
describe rheological phenomena.
The concept of a tensor was introduced into physics, and thus
into rheology, because it is useful; without it, the quantitative
description of many physical phenomena would be hopelessly clumsy
and tedious. Because of this usefulness, most of the literature on
viscoelastic behavior makes some use of tensor notation. This
literature will be inaccessible to a reader having no familiarity with
tensor quantities. Moreover, we will use tensors extensively in
several chapters of this book.
In the first section, we stated that rheology involves the relationship between deformation (strain) and force (stress) for a material.
It is in the quantitative description of the quantities strain and
stress that tensor notation is virtually indispensable. However, before demonstrating this, it will be useful to review briefly the
concept of a vector, as this is central to an understanding of tensors.

INTRODUCTION TO RHEOLOGY

23

1.7.1 Vectors

Certain physical quantities, such as force and velocity, are best


specified in terms of vectors, because a vector has a magnitude and
a direction. One example is the velocity vector, v. A vector can be
specified by giving its components, VI' v 2 , and v 3 ' referring to the
velocities in the three directions, Xl' x 2 , and x 3 Generally, we can
refer to a typical velocity component as Vi' where i can be 1, 2, or 3.
Note that while the magnitude is a physical attribute of a vector
that does not depend on the choice of a particular coordinate
system, the components of the vector do depend on the coordinate
system selected to describe the flow. There is a simple rule that tells
how to use the components of a vector in one coordinate system to
calculate the components of that vector in a second coordinate
system that is rotated with respect to the first.
If vectors are adequate to describe the velocity of a body and the
force acting on it, why are they not sufficient for describing rheological phenomena? The answer is that rheology deals not with motion
per se, but with deformation, and specifically with the relationship
between the deformation of a fluid element and the surface forces
exerted on this element by the surrounding fluid. Tensors are very
useful in specifying these two types of quantities, and the specific
tensors that are used to represent these quantities are the strain
tensor and the stress tensor.
1.7.2 What is a Tensor?

Like a vector, a tensor can be represented in terms of its components, and the values of these components depend on the choice
of the coordinate system used. Furthermore, there is a rule for
using the components of a tensor in one coordinate system to
calculate the components of that tensor in another coordinate
system, rotated with respect to the first. The existence of this rule
shows that a tensor has a basic physical significance that transcends
the arbitrary choice of the coordinate system. However, unlike a
vector, the physical significance of a tensor cannot be described in
terms of a directed line segment, i.e., in terms of a magnitude and a
direction.
In describing rheological behavior associated with a particular
type of deformation, we generally select a coordinate system that

24

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

gives the components a physical significance that is easily understood. For example, in describing the stretching of a rod ("uniaxial
extension") we take the x 1 direction to be the direction of stretching. Then the 0"11 component of the stress tensor is simply the
tensile stress in the sample.
Whereas a vector has three components, the tensors we will use
have nine. A typical component can be written using two indices,
for example O"ij' where each index can take on one of the values 1,
2, or 3 corresponding to the three coordinate directions. To present
the values of all the components of a tensor, matrix notation can be
used.
0"11
0" .. =
IJ

[ 0"21

(1-25)

0"31

Can any nine numbers form the components of a tensor? No. These
numbers have a specific mathematical significance, which we will
find particularly suited to the description of deformation and stress.
Specifically, these nine components contain all the information
necessary to transform one vector into another one that has a
certain prescribed relationship with the first. In mathematical language we say that the tensor "operates on" one vector to yield a
second vector, which contains information taken from both the
original vector and the tensor. For example, we will see that the
strain tensor, i.e. the nine components of the strain tensor, can
be used to operate on the components of the vector describing the
relative position of fluid particles within an undeformed fluid element, to yield the corresponding position vectors after deformation.
Likewise, the stress tensor can be used to operate on the unit
normal vector defining the orientation of a surface of a fluid
element to yield the surface force vector acting on that element.
Since the vector operated on in both cases is an arbitrarily selected
one, we see that the strain tensor actually contains a complete
description of the deformation that a fluid element undergoes
during some flow process, while tile stress tensor contains a complete description of the state of stress acting at a point in the fluid
at a particular time.
Our objective in this book is not to solve flow problems but only
to describe rheological phenomena. Thus, tensor calculus will not

INTRODUCTION TO RHEOLOGY

25

be required, and the reader need learn no new mathematics but


only the definitions of a few quantities.
With regard to notation, we will use a bold face symbol to
indicate that it is a vector. For example, the velocity vector will be
represented as v. The components of a vector will be indicated by
means of a subscript, for example, Vi' For the components of a
tensor we will use two subscripts. For example, the components of
the stress tensor will be represented by uij' In order to minimize the
number of new symbols and rules that need be learned, we will not
use dyadic notation or the Einstein summation convention. These
are methods of notation that simplify the writing of equations
involving tensors, and they are described in the book by Aris [1].
1.8 THE STRESS TENSOR

The deformations that occur in the processing and use of materials


are generally more complicated than simple extension and simple
shear and involve a combination of these two types of deformation.
For example, consider the deflection of a rubber tire under load or
the flow of a molten plastic into a mold. First, the deformation is
not uniform but varies from one place to another within the
material. It thus becomes a "field variable," i.e., a quantity that
varies from one point to another and is thus a function of position.
Secondly, the stress is not purely tensile, compressive or shear.
The quantitative specification of the forces acting on a solid
body as a result of contact with another body is straightforward;
one need only give the components of the force vector acting at the
interface. However, the specification of the forces acting on the
surface of a fluid element is less obvious, since the orientation of
the surface is arbitrary, i.e., it depends on how one defines a fluid
element. It would appear that in order to completely specify the
state of stress on a fluid element one would have to give the
components of the stress vector for every possible orientation of
the surface. Fortunately, this is not the case, and we will see that by
specifying only the components of the stress teDSQr, the state of
stress at a point in a fluid can be completely de,s.cribed.
For a given, arbitrary, choice of a fluid element, the surface stress
vector is t(o), where n is the unit normal vector for the surface and
specifies its orientation. The (n) superscript on the surface stress

26

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

> ____ x,

Figure 1-11. Cubical material element with a typical stress component.

vector indicates that the components of the vector depend on the


orientation of the surface. This vector does not at first seem to be a
useful tool for describing the state of stress at a point in a fluid
because of the arbitrariness of the choice of the surface orientation.
It is possible to show, however, that if the stress vectors acting on
each of three mutually perpendicular planes passing through a
point in a fluid are specified, the stress vectors for any other choice
of planes can be calculated by means of a simple transformation
rule [1]. It is convenient to let these planes be perpendicular to the
coordinate axes. Thus, the unit normal vector for a surface becomes
equal to one of the unit normal vectors for the coordinate system:

and the components of the force vector are given by:


t(e.l = (T ..
]

I]

where (Tij is the stress tensor. 2


To understand the physical significance of the nine components
of the stress tensor, consider the small cubical element of material
shown in Figure 1-11. The second subscript indicates the direction
of the force and corresponds to the coordinate axis direction. For
example, the stress component shown in Figure 1-11 acts in the Xl
2The components of the surface stress vector for any other surface whose orientation is
defined by the unit normal vector, n, can be determined as follows:

INTRODUCTION TO RHEOLOGY

/'------x,

27

-+---~al1

Figure 1-12. Several additional stress components.

direction, and the second subscript of this component is thus 1. The


first subscript indicates the face on which the component acts, and
this is specified by reference to the coordinate direction normal to
this face. Thus, the force shown acts on a face normal to the x 2
direction, and the first subscript of this component of the stress is
thus 2. The stress component shown is thus (T21'
To complete our definition of the components of (Tij' we need a
sign convention. In this book, we will use the convention generally
used in mechanics, although the reader should be aware that the
opposite convention is used by some rheologists [2-4]. We will take
the stress to be positive when it acts in the positive Xj direction, on
a face having the higher value of Xi' i.e., the face further from the
origin in the Xi direction. For example, the stress component shown
in Figure 1-11 is positive if the force acts in the direction of the
arrow. This is because it acts in the positive Xl direction on a face
having the higher value of x 2 Figure 1-12 shows several additional
components of the stress.
The set of nine components that is needed to specify completely
the state of stress at a point in a deformable material is an example
of a "second order tensor," and the members of the set are said to
be the "components" of the tensor. Since the components of the
stress tensor describe the state of stress at a point in the material,
the cubical element shown in" Figure 1-11 must be shrunk to an
infinitesimal size. Thus, the two force vectors shown in Figure 1-13
are acting in opposite directions at the same point. From Newton's
law of action and reaction, these two forces must be equal in
magnitude. They are thus both manifestations of the same compo-

28

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

nent of the stress tensor, 0"1l' and both have positive values if they
act in the directions indicated. Thus, according to our sign convention, a tensile stress has a positive value.

(111 . . .

-------I~(111

Figure 1-13. Equal, opposite forces at a point are represented by the same component of the
stress tensor.

The principle of conservation of angular momentum can be


applied to the infinitesimal material element we have been considering to show that the stress tensor has the following property:
(1-26)

Thus, any two components that have the same subscripts or indexes, but in reversed order, have the same value. A tensor that has
this property is said to be "symmetric." One result of this property
is that a symmetric tensor has only six independent components
rather than the nine that would be required to completely specify a
nonsymmetric tensor.
To make more concrete our discussion of stress, consider the
simple shearing deformation shown in Figure 1-14. There is a more
or less universal convention in describing this flow, and it is that the
direction of motion is Xl' while the velocity varies in the X 2
direction. To generate this deformation, a force is applied to the
upper plate in the direction shown by the arrow. In the ideal case,
(fully developed flow with no edge effects) this force generates a
uniform stress in the sample. Since the force is in the Xl direction
and acts on a face perpendicular to the X 2 direction, the stress
generated by the force F is 0"21' Obviously, this is a shear stress.
Due to the symmetry of the stress tensor this is equal to 0"12' We

Figure 1-14. Simple shear index convention.

INTRODUCTION TO RHEOLOGY

29

will use the symbol, (T, with no subscripts, to mean this component
of the stress tensor in simple shear. Thus for simple shear,

(1-27)
where A is the area of the sample in contact with the plates. The
other shear stress components are zero:

(1-28)
We can now describe, using matrix notation, the state of stress in a
material subjected to simple shear:

(1-29)

Another example of a test that is of practical interest in rheology


is simple or uniaxial extension. This test is illustrated in Figure 1-2.
If we let x 1 be the direction of the applied force, the stress
component resulting from this force will be (T11' which is a normal
stress. If it acts in the direction shown, it is a tensile stress. There
are no shear stress components in this case, and the components of
the stress tensor are as shown below:
(Tll
(T.
I]

o
(1-30)

There is an additional point regarding normal stresses that should


be mentioned here. While all materials are compressible to some
extent, in the case of molten plastics, quite high pressures are
required to produce a significant change in the volume of a sample.
For this reason, for many purposes these materials can be considered to be incompressible. Now consider what happens when we
subject an incompressible material to a compressive or tensile stress
that is equal in all directions, i.e., an isotropic stress or "pressure."

30

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The components of the stress tensor in this situation are as shown


below.

(1-31)
A stress of this type is said to be "isotropic." The minus signs result
from the fact that pressure is considered positive when it acts to
compress a material, whereas according to our sign convention
compressive stress is a negative quantity. In this situation, the
sample will be totally unaffected by the forces associated with this
pressure, i.e., it will not change its size or shape. Thus, such an
isotropic stress field is of no rheological significance. Only when
there are shear stresses acting, as in simple shear flow, or when the
normal stress components are different from each other will deformation occur in an incompressible material.
Another way of saying this is that if a rheological measurement
on an incompressible material is repeated at several different
ambient pressures, for example by placing the rheometer in a
hyperbaric chamber, the measurements at various pressures will
yield exactly the same values of all rheological properties.
This means that for an incompressible material a normal component of stress has no absolute rheological significance. Only
differences between two normal components are of rheological
significance. For example, in the case of simple shear, it is customary to describe the state of stress from a rheological point of view
by specifying the shear stress, a, and the "first and second normal
stress differences."

(1-32)
(1-33)
For Newtonian fluids these two quantities are zero in simple shear,
but in polymeric liquids they generally have nonzero values. One
manifestation of the first normal stress difference is observed when
a liquid is sheared by placing it between two flat parallel disks and
rotating one of the disks. It is found that an elastic liquid exerts a
normal thrust tending to separate the plates, while a Newtonian
fluid exerts no normal thrust on the plates.

INTRODUCTION TO RHEOLOGY

31

In the case of simple extension, there is only one rheologically


significant feature of the stress field, because there are no shear
stresses acting, and there is axial symmetry, so that a 22 = a 33 This
is the principal stretching stress, a E , defined in Equation 1-34.
(1-34)
Because the magnitude of a single normal stress component is
not rheologically significant by itself, it is customary in rheology to
define an "extra" or "viscous" stress as follows:
(1-35a)
(1-35b)
We see from Equation 1-31 that all the components of the extra
stress tensor are zero in a fluid that has been at rest for a long time.
However, for materials undergoing deformation, the stress is no
longer isotropic, and the "P" in the above equation no longer has a
direct relationship to the pressure as it is usually defined. However,
this will not be a problem for us, as we will work only with shear
stresses and normal stress differences. For these quantities, the
complete specification of P is not essential, and it matters not
whether we write our equations in terms of aij or Tij.
1.9 A STRAIN TENSOR FOR INFINITESIMAL DEFORMATIONS

In order to describe rheological phenomena in quantitative terms, it


is necessary to make use of a quantitative measure of deformation.
Such a measure is called strain, and we have defined in the previous
sections measures of strain suitable for specific types of deformation, namely, simple shear and simple extension. However, it is
advantageous to establish a general definition of strain that is valid
for all types of flow kinematics. Section 3.4 contains a discussion of
strain measures suitable for describing deformations of any size and
kinematics. For the present, we will define only the infinitesimal
strain tensor, which will be useful in the presentation of the theory
of linear viscoelasticity in Chapter 2.

32

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

B
A

(ORIGIN)

Figure 1-15. The position vector of a fluid particle (A) at times tl and
displacement vectors at time t2 for two neighboring fluid particles, A and B.

t 2,

and the

The simplest description of deformation of a fluid element is one


based on the behavior of the position vectors for the fluid particles
within this element. Figure 1-15 shows the position vector, xit 1),
for a typical particle "A" at a time, t 1, as well as the position vector,
xA(t Z )' for the same fluid particle at another time, t z . The use of
two times, or two states of the fluid element, is unavoidable in
discussing strain, as strain is a measure of the difference between
two states. We will take t1 to represent a time when the fluid is in a
reference configuration, with which the configuration at any other
time, t z , will be compared. Thus, we will say that a material
element is in a strained state at time t z if it undergoes deformation
in the time interval between t1 and t z .
To describe deformation it is necessary to look at two fluid
particles and the change that occurs in their relative positions
between times t1 and t z. In Figure 1-15, two such particles are
shown, with the position of the second particle with respect to the
first given by the vector i)x(t 1) at time t1 and by i)x(t z) at time t z.
The displacement vector, u(t z ) is the position of a particle at time t z
relative to its position at time t 1 The relative variation of the u(t z )
vectors for the two fluid particles obviously provides information
about the deformation of the fluid element containing these particles. For example, in a rigid displacement of the fluid element,
without rotation, as shown in Figure 1-16, the displacement vectors
for the two particles are the same, and this is true for any two

INTRODUCTION TO RHEOLOGY

33

t2

Figure 1-16. Displacement vectors for particles A and B, of a fluid element at time 12 and
after a rigid translation of the element.

particles in the fluid element. This suggests that the gradient of u


may be useful as a quantitative measure of strain. 3 This quantity
has nine components:

au;(t 2 )
ax/t l )

(1-36)

These nine quantItIes are the components of a tensor, and they


clearly contain all the information necessary to describe the relative
displacements of the two neighboring particles. This tensor has
some characteristics that make it unsuitable as a general measure of
strain, but if we limit our attention to flows in which this gradient is
always very small, a useful measure of strain can be defined as
follows:

(1-37)
where "fJtl' t 2) is the infinitesimal strain tensor at time t 2, relative
to the configuration of the same fluid element at time tl.4 Like the
3 In

some texts, this is called the "displacement gradient tensor," but we will reserve that
name for the tensor, Fij , introduced in Chapter 3.
4The definition used here is the one proposed by Ferry [5], but others will be found in the
literature, for example "Iij/Z"

34

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

stress, this is a symmetric tensor, because if the order of the indexes


i and j is reversed, the value of each component remains unchanged.
The "strain rate tensor," also called the "rate of deformation
tensor," is defined by Equation 1-38.
(1-38)
Since the components

Vi

of the velocity vector are given by:

(1-39)

duJdt

Vi =

Equation 1-38 can be expanded as shown in Equation 1-40.

aVoJ

IJ

avo

+ax. ax.J

1' .. = -

(1-40)

This definition of a rate of deformation is not limited to use in


describing deformations that are infinitesimal in total magnitude, as
its definition requires only a measure of strain valid during an
infinitesimal interval of time.
As an example of the use of the infinitesimal strain tensor, we
note that for simple shear of magnitude 1';

(1-41)
The variation with time of the components of the position vector
for simple shear are:

x t(t 2)

xt(tt) + x 2(tt)[ y(t2) - y(tt)]

(1-42a)

xz{t 2) = x 2(tt)

(1-42b)

x 3(tt)

(1-42c)

X3(t 2)

The only nonzero component of the displacement vector is u t:

INTRODUCTION TO RHEOLOGY

35

and the components of the gradient of the displacement vector are:

~l

(l-43)

Thus, the only nonzero components of the infinitesimal strain


tensor defined by Equation 1-37 are:
Y12 = Y21

==

(1-44)

and the only nonzero components of the rate of strain tensor are:
(l-45)
Likewise, for simple extension, it can be seen that in the limit of
infinitesimal deformation, the Hencky and linear strains become
equal to each other, and that both become equal to two times
au1laxl' All of the components of the deformation gradient tensor
for simple extension are shown below, where e == e(t 2 ) - e(t 1 ).

-e12

(1-46)

Thus, the components of the infinitesimal strain tensor are:

-e

(l-47)

For simple extension, the components of the strain rate tensor are:

-e

(1-48)

36

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1.10 THE NEWTONIAN FLUID

We can make use of the tensors defined above to write out the
general constitutive equation for the Newtonian fluid, which is valid
for low molecular weight, single phase liquids:
(1-49)
Whereas Equation 1-16 gives only the shear stress of a Newtonian
fluid undergoing steady simple shear, Equation 1-49 provides information about all the components of the stress tensor for any type
of deformation.
For example, we can use it to calculate all the stresses in steady
simple shear. The velocity distribution for this deformation can be
written as follows:
(1-50a)
(1-50b)
and the components of the rate of deformation tensor are:
)'

o
o

~l

(1-51)

Thus, the components of the extra stress tensor for a Newtonian


fluid are, from Equation 1-49,
T/)'

o
o

~l

(1-52)

We note that the normal components of the extra stress are all zero
in the Newtonian fluid. This means that:
(1-53)

INTRODUCTION TO RHEOLOGY

37

and that the normal stress differences are both zero.


(1-54)
It is important to remember that these equations for the components of the stress tensor are only valid for steady simple shear of a
Newtonian fluid.
Another example of the use of the constitutive equation for a
Newtonian fluid is the calculation of the stresses in a deformation
other than simple shear, namely simple extension. Combining
Equations 1-48 and 1-49 we have:

o
-10

(1-55)

For a Newtonian fluid, we have seen that it is possible to measure


the viscosity in a simple shear experiment and then use it in the
general constitutive equation 0-49) to determine all the components of the extra stress tensor for any type of flow. For non-Newtonian fluids this cannot be done, i.e., it is not possible to predict a
fluid's response to arbitrary deformations on the basis of rheological information obtained using only simple shear experiments.
1.11 THE BASIC EQUATIONS OF FLUID MECHANICS

The solving of the equations of fluid mechanics to predict specific


flows of molten plastics is beyond the scope of this book. This is the
procedure that is necessary, for example, to predict the detailed
flow pattern and velocity distribution for melt flow in a die or in a
mold, and there are several books that show how to do this [6-9].
However, in order to understand the role played by rheology in
solving such problems, we will simply present the basic equations of
continuum mechanics and make a few comments about them.
These basic equations are the continuity equation, which is derived
from the principle of conservation of mass, and Cauchy's equation,
which is a generalization of Newton's second law of motion,
(F = rna) for the case of a deformable material.

38

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

In deriving these equations, which govern the mechanical behavior of deformable materials, it is assumed that every point in space
(specified by Xl' X 2 , x 3 ) within a material has associated with it
values for all the components of the velocity vector, the strain
tensor and the stress tensor. It is further assumed that these "field
variables" are continuous functions of position. In reality, we know
that this is not precisely true, because at a submicroscopic level we
find mostly empty space punctuated by regions of concentrated
mass, which are the atoms of the molecules of which the material is
composed.
However, the scale of our macroscopic observation and of practical flow phenomena is so large compared to the scale at which this
inhomogeneity appears, that for all practical purposes we can
assume the material to be a "continuum" and make use of the
strain, velocity and stress fields to describe the deformation and
flow properties of the material. This is called the "continuum
assumption." Thus, we assume that the collection of molecules that
make up the material of interest can be represented, for purposes
of describing its flow and deformation, by a mathematical continuum. The equations based on this assumption are thus called the
equations of continuum mechanics.
1.11.1 The Continuity Equation

The principle of conservation of mass, when applied to a deformable continuum, tells us that the net flow of mass into any fixed
volume element contained within the continuum during some interval of time is equal to the accumulation of mass within the element
during that time interval. This principle is summarized by Equation
1-56.
[NET FLOW IN]

[FLOW IN] - [FLOW OUT]


[ACCUMULATION]

( 1-56)

For an incompressible material, the accumulation of mass within a


volume element is zero. We will not present a derivation here, but
in this case it can be shown that the principle of conservation of

INTRODUCTION TO RHEOLOGY

39

mass can be written as shown in Equation 1-57.


aV l

aV 2

aV 3

aX l

aX 2

aX 3

-+-+-=0

(1-57)

This is the "continuity equation" for an incompressible material.


1.11.2 Cauchy's Equation

The second basic equation of continuum mechanics is based on


Newton's second law of motion (F = rna). This basic principle
describes the relationship between the applied force and the change
of momentum of a rigid body. It was later put into a form suitable
for describing the deformation of a continuum in response to
deforming stresses, and the result is Cauchy's equation.
(1-58)

The quantity on the left is the density times the "substantial


derivative," which is defined as follows:
(1-59)

The substantial derivative is the quantity that tells how the velocity
of a fluid element changes with time as it moves through the flow
field.
Equations 1-58 and 1-59 are vector equations, and since i can
have anyone of the three values 1, 2, or 3, each can also be written
as a set of three scalar equations.
Making use of the definition of the extra stress (Eq. 1-35),
Cauchy's equation can be rewritten in a form involving the pressure.
(1-60)

The left hand side describes the acceleration of a material element,

40

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

while the right hand side accounts for the applied forces. The first
term on the right represents the force due to the earth's gravitational field, which is a "body force" and acts directly on the mass of
the material element. The second term describes the net force on a
material element due to pressure gradients, and the last three terms
involve the extra stress tensor and result from the deformation of
the fluid. It is through these last three terms that the rheological
properties of the material become involved in its mechanical
behavior.
For molten polymers, the very high resistance to deformation
resulting from the strong interactions between molecules makes the
last three terms very large compared to the acceleration term on
the left hand side. For this reason, in flow situations involving
molten plastics, the left hand side can be neglected. This is called
the "creeping flow" approximation.
Taking account of the fact that Cauchy's equation represents
three scalar equations, and adding in the continuity equation, we
have four equations. The number of dependent variables is much
greater than this, however, as it includes the three components of
the velocity vector plus the six independent components of the
stress tensor. Thus, no problem in continuum mechanics can be
solved without an additional equation. This additional equation is
the rheological constitutive equation, which relates the components
of the stress tensor to the deformation.
1.11.3 The Navier-Stokes Equation

The constitutive equation for a Newtonian fluid is given by Equation 1-49. When this is combined with Cauchy's equation 0-60) and
the continuity equation is used to simplify the terms, the result is
the Navier-Stokes equation, which governs the flow of Newtonian
fluids. If the viscosity is assumed to be a constant, i.e., independent
of position, the Navier-Stokes equation can be written as follows:

Dv
I

p-

Dt

ah
ax;

ap
( a2 v
a2 v
a2 v )
+ 7] ++(1-61)
ax;
axfax~
ax~

= -pg- - -

Note that this is a vector equation and that by setting i equal to 1, 2


and 3, we can rewrite it as three scalar equations.

INTRODUCTION TO RHEOLOGY

41

In the analysis of flows involving Newtonian fluids, a dimensionless group of central importance is the Reynolds number. This is
defined as follows:
DVp

Re=-71

(1-62)

where D and V are a characteristic length and velocity respectively.


This is a measure of the relative importance of viscous and inertial
(acceleration) effects. In other words, it is a measure of the relative
importance of the third term on the right of Equation 1-61 as
compared with the term on the left. For fluids with very high
viscosity, the Reynolds number is generally very low, and the inertia
or acceleration term on the left hand side of Equation 1-61 can be
neglected. This is often referred to as the "creeping flow" approximation. While molten polymers are not Newtonian, and Equation
1-61 is therefore not appropriate, it is generally accepted that for
these materials, the creeping flow approximation is valid. We will
make use of this concept in Chapter 8 when we discuss the analysis
of entrance flows.
REFERENCES
1. R. Aris, Vectors, Tensors and the Basic Equations of Fluid Mechanics, PrenticeHall, Inc., Englewood Cliffs, N.J., 1962.
2. R. B. Bird, W. E. Stewart and E. N. Lightfoot, Transport Phenomena, John
Wiley & Sons, New York, 1960.
3. R. B. Bird, R. C. Armstrong and O. Hassager, Dynamics of Polymeric Liquids:
Volume 1, Fluid Mechanics, Second Edition, John Wiley & Sons, New York,
1987.
4. R. B. Bird, O. Hassager, R. C. Armstrong and C. F. Curtiss, Dynamics of
Polymeric Liquids: Volume II, Kinetic Theory, Second Edition, John Wiley &
Sons, New York, 1987.
5. J. D. Ferry, Viscoelasticity of Polymers, John Wiley & Sons, N.Y., 1980.
6. S. Middleman, Fundamentals of Polymer Processing, McGraw-Hill, New York,
1979.
7. Z. Tadmor and C. G. Gogos, Principles of Polymer Processing, John Wiley &
Sons, New York, 1979.
8. M. J. Crochet, A. R. Davies and K. Walters, Numerical Simulation of NonNewtonian Flow, Elsevier Science Publishing Co., New York, 1984.
9. R. Keunings, "Simulation of Viscoelastic Fluid Flow", in Fundamentals of
Computer Modeling for Polymer Processing, C. L. Tucker II, Ed., Hanser Publishers, Munich, 1988.

Chapter 2
Linear Viscoelasticity
2.1 INTRODUCTION

The simplest type of viscoelastic behavior is linear viscoelasticity.


This type of behavior is observed when the deformation is sufficiently mild that the molecules of a polymeric material are disturbed from their equilibrium configuration and entanglement state
to a negligible extent. Obviously, very small deformations would be
in this category. This might be a deformation in which the total
strain was very small, or the early stages of a larger deformation.
For melts, which have a fading memory and can flow, linear
behavior is also observed when a deformation occurs very slowly, as
in steady simple shear at very low shear rates. This is because
relaxation processes due to Brownian motion are always acting to
return the molecules to their equilibrium state, and if the deformation is tending to take them away from this state only very slowly,
this relaxation mechanism has plenty of time to "keep up" with this
process, with the net result that no significant deviation from
equilibrium occurs. One manifestation of this is that at very low
shear rates, the viscosity of a polymeric liquid becomes independent
of shear rate.
Since the deformations that occur in plastics processing are
neither very small nor very slow, one might wonder how the theory
of linear viscoelasticity could be put to practical use. It is clearly of
little use in process modelling. In fact, its principal utility is as a
method for characterizing the molecules in their equilibrium state.
This is useful in the comparison of different resins, for example for
resin quality control. In the case of linear polymers, it is sometimes
possible to correlate linear viscoelastic properties with molecular
weight or molecular weight distribution. The concept of molecular
42

LINEAR VISCOELASTICITY

43

weight distribution is described in Appendix B, and quantitative


definitions of the various averages are given there as well.
In this chapter, we present the theory of linear viscoelasticity
along with formulas that are useful for the treatment of linear
viscoelastic data. We then describe the linear behavior of molten
polymers and the molecular mechanisms underlying this behavior.
A more thorough treatment of this subject can be found in the
excellent monographs by Ferry [1], on polymer viscoelasticity, and
by Tschoegl [2], on the mathematical description of viscoelastic
behavior.
2.2 THE RElAXATION MODULUS

In Section 1.5.3 we mentioned an experiment in which the Maxwell


element shown in Figure 1-8 was subjected to a sudden elongation,
and the force was calculated as a function of time. The analogous
rheological experiment is one in which a sample is suddenly deformed at time t = 0, and the resulting stress is measured as
a function of time. This is called a "stress relaxation" experiment.
In reporting the results of such an experiment, it is customary to
divide the stress by the magnitude of the strain that is introduced to
start the experiment. In the case of a sudden shear strain of
magnitude 'Yo, the quantity reported is the "shear relaxation modulus," G(t , 'Yo)'

G(t, 'Yo) == IT(t)/'Yo

(2-1)

In the case of an extensional strain of magnitude So' the quantity


reported is the "tensile relaxation modulus," E(t, so), which is the
net tensile stress, lTE' divided by so'
(2-2)
In general, the relaxation moduli defined above are functions of
the strain magnitude. When the strain is very small, however, they
are independent of the strain. In this case the stress at any particu-

44

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

lar value of t is proportional to the strain.


(2-3)

u(t) = G(tho

It is because of linear relationships such as these that we call this


small-strain behavior "linear" viscoelasticity.
2.3 THE BOLTZMANN SUPERPOSITION PRINCIPLE

It is useful to have a general equation that describes all types of


linear viscoelastic behavior. We will derive this equation starting
from the assumption that the relaxation modulus is independent of
strain amplitude, i.e., from Equation 2-3. Consider a sequence of
small shear strains as shown in Figure 2-1. The shear stress resulting from the strain that occurs at time t1 will be:

(2-4)
To calculate the stress resulting from the strain introduced at time
t 2 , we assume that the incremental response of the material to this
second step strain is independent of the strain introduced at time
t 1. Thus, we can simply add on or superpose the stress resulting

8)'(t3 )

II:
I-

en

II:

~
en

8)'(t2 )

:I:

f
8)'(t,)
o

t,
TIME

Figure 2-1. Sequence of step strains.

8)'(t4)

LINEAR VISCOELASTICITY

from the strain at time

45

as follows:

t2

For any combination of N small strains, we can continue to add on


the contributions to the stress, and in general:
N

O"(t)

L G(t -

tJoy(t;)

(2-6)

i= 1

For a smooth strain history not consisting of finite steps, we can


make use of the definition of the integral to show that:

O"(t)

G(t - t') dy(t')

(2-7)

-00

Noting that the strain that occurs during the time interval dt' is
simply yet') dt', this can also be written as:

O"(t)

G(t - t')y(t') dt'

(2-8)

-00

The use of the lower limit of minus infinity is a mathematical


convenience; it implies that to calculate the stress at time t in the
most general case one must know the strain history infinitely far
into the past, i.e., at all times t' prior to t. In practice this is not
necessary. We generally start an experiment at some time t = 0
when the material is in a stress free state. In this case, 0"(0) = 0,
and:

O"(t)

{G(t - t') dy(t')


o

(2-9)

The above equations apply only to shearing deformations, but


they can be generalized for any type of deformation by using
another feature of linear viscoelasticity. This is that the relaxation
process is independent not only of the magnitude of the strain but
also of the type (kinematics) of the deformation. Thus, we can
replace the shear strain by the strain tensor for infinitesimal strain,
and we can replace the shear stress by the stress tensor to obtain

46

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the following two alternative forms of the "Boltzmann superposition principle":

T;/t) = {

G(t - t') dY;j(t ')

(2-10)

G(t - tl)Yij(t') dt '

(2-11)

-00

T;j(t) = {
-00

The tensorial forms of the Boltzmann superposition principle


(Equations 2-10 and 2-11) can be used, for example, to see how
extensional flow properties are related to the shear relaxation
modulus. Consider the extensional flow experiment described in
Section 1.4.1. For an incompressible material undergoing a step
strain deformation, of magnitude EO' the components of the infinitesimal strain tensor are, from Equation 1-47:

o
(2-12)

Thus, the components of the extra stress tensor can be determined


from Equation 2-10, and these are shown below.

o
(2-13)

The net stretching stress is thus as shown in Equation 2-14.


(2-14)

In other words, for an incompressible material exhibiting linear


viscoelastic behavior, Young's modulus is three times the shear
modulus.
O"E(t)/E O = E(t) = 3G(t)

(2-15)

LINEAR VISCOELASTICITY

47

Using the Boltzmann superposition principle, it is possible to


calculate the stress components resulting from any type of deformation, as long as that deformation is sufficiently small or slow that
linear behavior is exhibited. We have already seen how the linear
relaxation moduli for shear G(t) and extension E(t) are related.
Another example is steady simple shear. Here, we assume the
components of the strain rate tensor are given at all times t' prior
to the present time, t, by:
'Y

o
o

~l

(2-16)

where y is a constant. In practice, it is not necessary that this flow


pattern have been maintained at all past times. Because elastic
liquids have a fading memory, the actual time required for the
stresses to become steady when the flow is started up is usually on
the order of seconds, or minutes at the most. From Equation 2-11
we can determine the components of the extra stress tensor for
steady simple shear. Only two of these are nonzero:
7"12

7"21

= y{

G(f - f') dt'

(2-17)

-00

Or, letting s ==

f - f',

1 G( s) ds

00

(2-18)

Since the definition of the viscosity is: Tl == a / y, the low-shear rate


limiting viscosity, Tlo, which is the value predicted by the theory of
linear viscoelasticity, is:
Tlo

1 G(s) ds
00

(2-19)

The diagonal components of the extra stress tensor are all zero, so
N1 = N2 = O. In fact, by inspection of Equations 2-11 and 2-16 one
can see that the normal stress differences will be zero in any simple

48

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

shear deformation, whether or not it is steady in time, as long as the


linear theory is valid, i.e., as long as the deformation is sufficiently
small or slow.

2.4 RElAXATION MODULUS OF MOLTEN POLYMERS

Figure 2-2 is a plot of G(t) for a typical crosslinked elastomer. The


general features of the curve can be understood qualitatively by
reference to Figure 1-7, which shows the mechanical assembly
called a Kelvin or Voigt body. Let us take the spring to represent
the Brownian motion spring mechanism of polymer elasticity and
the dashpot to represent friction between molecules. Clearly there
is no way that this assembly can accommodate the sudden strain
imposed at the beginning of the stress relaxation experiment, as the
shear rate is infinite and the dashpot force would also be infinite. In
fact, the rubber deforms initially by a molecular mechanism that
has no analogy in the Voigt body. This is the mechanism of
chemical bond distortion and is the origin of the "glassy" behavior
observed in most materials at sufficiently short times and or sufficiently low temperatures. In a very qualitative sense it is as if an
extremely stiff spring were placed in series with the Voigt body. The
logarithmic time scale is used in Figure 2-2 to expand the short time

log(time)
Figure 2-2. Relaxation modulus for elastomer.

LINEAR VISCOELASTICITY

49

109(t)

Figure 2-3. Relaxation modulus for three samples of a typical linear polymer. A is monodisperse with M < M6 B is monodisperse with M Me, and C is polydisperse with Mw Me.

glassy behavior and show that there is a very short initial time
period during which this primarily elastic, glassy mechanism
operates.
Once the initial deformation is accommodated by the glassy
mechanism, changes in molecular conformation become possible via
Brownian motion, and the stress decays. Since this is a crosslinked
material, the modulus will approach an equilibrium value, Ge , at
long times.
Turning now to uncross linked melts, we find behavior similar to
that of a rubber at very short times, but at longer times, the melt
flows instead of approaching an equilibrium configuration. Figure
2-3 shows typical G(t) curves, again using logarithmic scales, for
three samples of a linear polymer. Two samples, A and B, have
narrow molecular weight distributions. 1 One of these, B, has a
molecular weight that is sufficiently high that there is a significant
degree of entanglement. Sample A has a molecular weight below
that at which there is a significant level of entanglement, while C
has a broad molecular weight distribution.
We note that for all three materials there is a very short time
zone in which glassy behavior is observed. All three materials
ISee Appendix B for an explanation of the molecular weight distribution.

50

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

exhibit the same behavior in this region as only very small scale
molecular motions are involved, and the molecular weight is not
important. As in the case of the crosslinked rubber, relaxation by
the Brownian motion mechanism cannot occur at this very short
time scale, and the material is very stiff, i.e. it has a high initial
value of G(t). Glassy behavior is observed over ever longer time
scales as the temperature is lowered, and below the glass transition
temperature (See Appendix D) this is the only mode of behavior
observed. 2
Beyond the glassy zone, we have a transition zone where Brownian motion relaxation processes begin to be active. For the low
molecular weight material (A) we then move directly to a "terminal
zone" where G(t) falls off toward zero, as must happen eventually
for an uncrosslinked polymer. For the high molecular weight,
narrow distribution melt, B, the transition zone is followed not by a
terminal zone but by a plateau zone in which the modulus is nearly
constant. The modulus in this region is called the plateau modulus,
and has the symbol Gt. At longer times flow occurs, and the G(t)
curve moves into a "terminal zone" where the modulus relaxes,
eventually to zero at sufficiently long times. Except at exceedingly
high molecular weights, a true plateau is not achieved, and there is
a mild decrease in G(t) in the "plateau region." Ferry [1, p. 372]
gives empirical formulas for estimating a value of Gt in this case.
The existence of a plateau implies that there are two relaxation
mechanisms, each with a different set of relaxation times. There is
one type of relaxation that occurs at very short times, and this
governs the behavior in the transition zone. And there is another
type that occurs over a range of longer relaxation times, which
governs the terminal zone behavior. When the times over which
these two relaxation mechanisms operate are widely separated,
there is an intervening range of times over which little relaxation
occurs, and the melt appears to behave like the crosslinked material
shown in Figure 2-2.
This rubbery behavior implies the presence of strong interactions
between molecules that can mimic the effects of chemical crosslinks
over a certain period of time. These strong interactions between

2In fact, certain very restrictive types of molecular motion can occur even then, and this gives
rise to a small stress relaxation effect [1, Ch. 15].

LINEAR VISCOELASTICITY

51

molecules are usually called "entanglements," as they are clearly


different from the simple hydrodynamic frictional interactions that
occur in the lower molecular weight liquids. Because these strong
interactions result in behavior similar to that exhibited by a rubber,
which has a permanent network, they are sometimes said to comprise a "temporary network." By analogy with the behavior of a
crosslinked rubber, the plateau modulus has been used to define an
"average molecular weight between entanglements," Me:
GO

pRT

=--

Me

(2-20)

The curves shown in Figure 2-3 for samples A and B are typical
of linear polymers having a narrow molecular weight distribution.
Note that the plateau modulus is a characteristic of the polymer
and is independent of the molecular weight. The curve for sample
C shows the type of behavior observed for polymers with a broad
molecular weight distribution. There is neither a well-defined
plateau zone nor a sharp transition to a well-defined terminal zone.
It is not possible, therefore, to determine a reliable value for the
plateau modulus. Long chain branching has a similar effect on
relaxation modulus behavior.
It is important to take note of the practical importance of the
behavior in the various zones of behavior of polymeric materials. At
temperatures above the glass transition temperature, glassy behavior is rarely if ever observed in deformations of practical interest. In
melt processing operations, it is the plateau and terminal zones that
are most important. For this reason, we will make no further
reference to behavior in the glassy zone.
2.5 EMPIRICAL EQUATIONS FOR THE RELAXATION MODULUS

The theory of linear viscoelasticity presented in the previous section


shows that if the relaxation modulus, G(t), is determined in any
experiment, then the linear response to any type of deformation
can, in principle, be predicted. However, it is almost essential when
carrying out such calculations to have G(t) given as an equation
rather than a graph or table of data. It is convenient to have a
general form for G(t) that contains sufficient parameters to fit

52

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

experimental data for a wide range of polymeric liquids. Several


such equations are presented in this section.
2.5.1 The Generalized Maxwell Model

The most popular approach to establishing such a functional form


is based on the use of the Maxwell element analogy described in
Section 1.5.3. If we take the spring constant, K e, to be analogous to
the initial shear modulus, Go, of the polymeric liquid, and the time
constant, Kv!Ke' to be analogous to the relaxation time of the
liquid, A, then in a step shear strain experiment:

O'(t)

Go'Yo[exp( -t/A)]

(2-21)

The shear relaxation modulus is thus:

G(t) = Go[exp( -t/A)]

(2-22)

and from Equation 2-11 the linear constitutive equation is

Tij(t) =

Go{exp[ -(t - t')/A]}Yij(t') dt'

(2-23)

-00

This is called the Maxwell model.


As was noted in Chapter 1, actual relaxation processes cannot be
described by a single exponential function. However, great flexibility can be obtained by use of the "generalized Maxwell model,"
which is the rheological constitutive equation analogous to the
mechanical assembly shown in Figure 2-4. The forces in the various
elements are additive, and the generalized Maxwell model is:

Tij(t)

EGk{exp[ -(t - t')/AkJ}Yij(t') dt'

(2-24)

-00

where G k and Ak are the initial modulus and relaxation time


corresponding to each Maxwell element. The relaxation modulus is:
N

G(t)

E Gi[exp( -t/AJJ
i= 1

(2-25)

LINEAR VISCOELASTICITY

53

Figure 2-4. Mechanical analog of the generalized Maxwell model.

By use of a sufficient number of elements, this equation can be


made to describe almost any experimental G(t) behavior. Between
five and ten elements or G i - Ai pairs are usually sufficient to fit
experimental data reasonably well. Such a set of values is called a
"discrete spectrum" of the material. Methods for determining a
discrete spectrum for a material are discussed in Section 2.9.
If G(t) is approximated by use of the generalized Maxwell model
with a finite number of elements, the behavior of G(t) at sufficiently long times will be dominated by the G i - Ai pair having the
largest value of Ai' and in the terminal zone G(t) will therefore
decrease exponentially with time. This implies that on a plot of
10g(G) versus log(t), such as Figure 2-3, the curve will approach a
line having a slope of minus one in the terminal zone. This largest
Ai is called the "terminal relaxation time."
2.5.2 Power Laws and an Exponential Function

For practical purposes, the generalized Maxwell model continues to


be the most used equation to represent relaxation modulus data.
However, many terms are often required, and this means that many
parameters are involved. For polymers having a broad molecular
weight distribution or those having a high degree of long chain
branching the modulus and spectrum are rather smooth and can
often be fitted by simpler equations having only a few empirical

54

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

constants. Larson [3,4] noticed that the relaxation modulus for


linear polymers with a broad molecular weight distribution and for
highly branched polymers can often be approximated by a power
law of the type shown below.

G{t) = ct- m

(2-26)

This empirical approximation is only useful in the transition and


terminal zones and has quite unrealistic behavior at short times.
Wagner [5] found that for a low density (highly branched) polyethylene the following modified exponential function gave a good fit
to experimental modulus data.

G{t) = {1Jo/2A)VA/t [exp (- Vt/A )]

(2-27)

In spite of their simplicity, little use seems to have been made of


the above two equations.
2.6 THE RELAXATION SPECTRUM

In the computation of one linear viscoelastic function from another,


it is sometimes convenient to make use of a continuous spectrum.
This can be defined by letting the number of elements in the
generalized Maxwell model increase without limit so that G(t) can
be represented in terms of a continuous function, F(A), such that
FdA is the contribution to the modulus from relaxation times
between A and A + dA. The relaxation modulus is related to the
spectrum as follows.

G{t)

[Co F{A)[exp( -t/A)] dA

(2-28)

However, a logarithmic time scale is normally used for the spectrum, and the continuous spectrum, H(A) is used in place of F(A),
where H = FA, and Hd In(A) = FdA. The relaxation modulus is
related to H(A) as follows.
G ( t)

{"

H ( A)[ exp ( - t / A)] d (In A)

-00

(2-29)

LINEAR VISCOELASTICITY

55

Ferry [1] has reviewed methods for determining the spectrum


from experimental data and for using it to calculate other material
functions. However, G(t) data are never available for the entire
range of times from zero to infinity. Furthermore, because of the
form of the relationship, a unique spectrum cannot be inferred
from experimental modulus data. This is an example of a "mathematically ill-posed problem," and special techniques may be required to avoid computational difficulties [6, 7].
Wissbrun [8] has proposed a spectrum function that is expressed
in the form shown in Equation 2-30.
(2-30)
where Ao is a characteristic time and b is the negative of the slope
of the plot of log 7J versus log y. This form was inspired by empirical observations of linear behavior.
2.7 CREEP AND CREEP RECOVERY: THE COMPLIANCE

In a creep test, a material initially in its equilibrium state is


subjected to a constant shear stress, a, at time t = O. The shear
strain, yet), is then monitored as a function of time. This is the
preferred experimental technique when it is desired to study linear
viscoelastic behavior over the broadest possible range of time. The
results are presented in terms of a material function called the
"shear creep compliance" defined as follows:

let) == y(t)ja

(2-31)

In the case of linear behavior, the creep compliance is independent


of a. For a melt, the shear rate will eventually become constant and
equal to the shear stress divided by the viscosity. If the intercept on
the I axis is defined as the "steady state compliance" (JJ),3 the
3 Ferry [1] uses the symbol leo for the steady state compliance, by analogy with the
"equilibrium" compliance of a crosslinked material. However, for a melt, there is no
equilibrium compliance.

56

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

.......

:::;ui

o
z

:J
a..

::E

oo

a..

w
w

a:

o JO
a:
w
J:

(J)

TIME(t)

Figure 2-5. Sketch showing a typical creep compliance curve for a melt.

compliance at long times is given by:

l(t) =ll

+ t/110

(2-32)

A typical creep compliance curve is sketched in Figure 2-5.


Using the Boltzmann principle, it can be shown that:

(2-33)

In terms of the parameters of the generalized Maxwell model:


LGjA~

(LGjAjt

(2-34)

The steady state compliance is a useful property for material


characterization. For example, for a high molecular weight, linear,

LINEAR VISCOELASTICITY

57

narrow distribution polymer the terminal relaxation time is given


approximately by:
(2-35)

The steady state compliance has been found to be independent


of the average molecular weight but strongly affected by the molecular weight distribution. Kurata [9] has reviewed a number of
equations that have been proposed to relate JJ to various moments
of the molecular weight distribution.
If at a time to' after the start of a creep experiment, the shear
stress is suddenly removed, for example by removing a weight from
a pulley mechanism, the material will spring back or recoil in a
direction opposite to that of the original applied force. The amount
of recoil or recovered shear strain is a function of to and of the time
(t - to) that has elapsed since the elimination of the shear stress:
(2-36)
If the stress is removed only after steady state has been achieved,
then the recoil is no longer a function of to, and it is convenient to
"reset the clock" for the observation of recoil at time to, i.e., to set
to = O. Then we have:

YrCt) == yeO) - yet)

(2-37)

The ultimate recoil, or "recoverable shear," when the material has


come to rest, is:
(2-38)

The recoil function or recoverable compliance, R(t), is defined as


follows:
R(t) == yJu

(2-39)

The symbol Jr(t) is also used for this material function. Making use

58

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

of the Boltzmann superposition principle, it can be shown that:

R{t)

let) - t/7]o

(2-40)

The ultimate value of the recoil function is therefore equal to the


steady state compliance:
lim [R(t)]

t->oo

IJ

(2-41)
(2-42)

Analogous to the shear creep and creep recovery functions described above are the extensional creep and creep recovery functions. Rather than applying a constant shear stress, we apply a
constant net tensile stress, O"E' starting at time t = O. The tensile
creep compliance is the Hencky strain divided by O"E.

D{t) == E{t)/O"E

(2-43)

By use of the Boltzmann superposition principle, it can be shown


that D(t) = l(t) /3. Thus, at long times, when D(t) becomes linear
with time, we have:

D(t) = IJ/3

+ t/37]0

(2-44)

Figure 2-6 shows sketches of shear creep compliance curves for


the three samples of linear polymer whose relaxation moduli are
shown in Figure 2-3. At very short times glassy behavior is observed,
and the compliance is constant. At times sufficient to allow motion
within a molecule, all three samples exhibit a transition zone. For
the low molecular weight sample, A, this transition leads directly to
a steady flow or terminal region, while for sample B, there is a well
defined plateau and then a sharp transition to a terminal zone. The
value of the compliance in this plateau zone is called l~. The broad
molecular weight sample, C, has neither a clearly defined plateau
nor a sharp transition to the terminal zone.
Figure 2-7 shows the recoverable compliance or recoil function,
R(t), for samples A, Band C. As shown by Equation 2-41, the long
time limiting value of this function is equal to the steady state

LINEAR VISCOELASTICITY

59

log(t)

Figure 2-6. Creep compliance curves for the three samples of a typical linear polymer. A is
monodisperse with M < Me; B is monodisperse with M Me, and C is polydisperse with
Mw Me
~-------------------------------------------,J.0
s

log(t)

Figure 2-7. Recoil functions for three samples of a linear polymer. A is monodisperse with
M < Me; B is monodisperse with M Me, and C is polydisperse with Mw Me.

60

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

compliance. Whereas lJ is only somewhat larger than l~ for the


monodisperse sample, it can be much greater than l~ for polydisperse materials, i.e., when the molecular weight distribution is
broad.
2.8 SMALL AMPLITUDE OSCILLATORY SHEAR

While it is convenient to use the shear relaxation modulus, G(t), to


introduce basic concepts, the experiment that has been most widely
used to determine the linear viscoelastic properties of polymeric
liquids is small amplitude oscillatory shear. In this experiment, a
thin sample of material is subjected to a simple shearing deformation such that the shear strain as a function of time is given by:

y{t)

Yo sin{wt)

(2-45)

where Yo is the strain amplitude and w is the frequency. The stress


is then measured as a function of time. By differentiating, we find
that the shear rate as a function of time is given by Equation 2-46.

y{t) = YoW cos{wt) = Yo cos{wt)

(2-46)

where Yo is the shear rate amplitude.


If Yo is sufficiently small, the stress can be calculated by use of
the Boltzmann superposition principle. Thus, by substituting Equation 2-46 into Equation 2-8, it can be shown that the stress is
sinusoidal in time and has the same frequency as the strain.

(T{t)

(To sin{wt + 5)

(2-47)

where (To is the stress amplitude and 5 is a phase shift, which is


called the "mechanical loss angle." Furthermore, we find that the
amplitude ratio, Gd (equal to (To/YO>, and the loss angle, 5, are
functions of frequency but are independent of the strain amplitude,
as long as Yo is sufficiently small that the Boltzmann superposition
principle is valid.
The Boltzmann superposition principle implies that the two functions of frequency, 5 and Gd (= (To/Yo), contain the same information. While there is no simple conversion formula, Booij and
Thoone [10] have presented the following formula, which they have

LINEAR VISCOELASTICITY

61

90
80
70
'Ul
Q)
~

C)
Q)

60

:8- 50
w
...J
(!)

z
40
en
en
0

...J

30
20
10

10

10~1

101

--

10 2

FREQUENCY (rad/s)
Figure 2-8. Mechanical loss angle as a function of frequency for several polymeric liquids.
1: PS, M = 2 X 10 5 (narrow MWD), T = 160C; 2: PS, M = 3.7 X 10 4 (narrow MWD),
T = 160C; 3: HDPE, Mw = lOS, Mw/Mn = 7.5, T = 180C; 4: HDPE, Mw = 1.5 X 10 6,
T = 190C; 5: PVC, 130C; 6: PVC, 205C. Adapted from Ref. 11. Copyright 1982 by
Steinkopff Verlag. Reprinted by permission.

found to be valid for a number of materials:


_

7T

o( w) - 2

[dd (In(InG
1
w)
d)

(2-48)

Booij and Palmen [11] note that o(w) is a sensitive measure


of differences between materials, and this is illustrated in
Figure 2-8.
2.8.1 The Complex Modulus and the Complex Viscosity

The results of an oscillatory shear experiment can be presented by


means of plots of the amplitude ratio and the phase shift as
functions of frequency. However, neither of these quantities has

62

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

any direct relationship with the material functions usually used to


describe viscoelastic behavior. For this reason, it is customary to
make use of a trigonometric identity to write Equation 2-47 in the
form shown in Equation 2-49.

a(t)

Yo[G'(w)sin(wt) + G"{w)cos(wt)]

(2-49)

where G'(w) is called the "storage modulus" and G"(w) is called


the "loss modulus." These two quantities can be easily calculated
from the amplitude ratio and the phase shift as follows:
G'

Gd cos(8)

(2-50)

Gil

Gd sin(8)

(2-51)

To understand the physical significance of the storage and loss


moduli, it is useful to examine the behavior of a Newtonian (linear,
purely viscous) fluid and a Hookean (linear, purely elastic) solid
when subjected to oscillatory shear. For the linear solid having a
shear modulus of G, we have:
Gy

(2-52)

YoG[sin(wt)]

(2-53)

so for oscillatory shear:

Thus, G' = G and Gil = O. The loss angle is zero in this case.
For the linear (Newtonian) fluid, we have:
a = T/Y

(2-54)

so for oscillatory shear:

(2-55)
In this case, G' = 0, Gil = T/W, and the loss angle is 7T /2 or 90,
since sin(wt + 7T /2) = cos(wt).
Thus, in a purely elastic material, where there is no viscous
dissipation, there is no phase shift, and the loss modulus is zero,
while for a purely viscous liquid, where there is no energy storage,

LINEAR VISCOELASTICITY

63

the storage modulus is zero, and the loss angle is 'TT /2. This
suggests that G'(w) and G"(w) can be interpreted in terms of
energy storage and loss respectively. In fact, it can easily be shown
that the energy dissipation per cycle per unit volume in oscillatory
shear is proportional to G". The work input per unit volume per
cycle, W, is:
(2-56)
Substituting from Equations 2-46, 2-47 and 2-51 and integrating, we
find that:
(2-57)
The energy dissipation is thus directly proportional to the loss
modulus.
It is sometimes convenient to think of G'(w) and G"(w) as the
real and imaginary components, respectively, of a "complex modulus" defined as follows:

G*{w) == G'(w) + iG"(w)

(2-58)

The amplitude ratio, Gd , is thus the magnitude of G*:


(2-59)
An alternative representation of the results of a small amplitude
oscillatory shear test makes use of two material functions, 11'( w),
the "dynamic viscosity," and 11"(W), both of which have units of
viscosity:

o-(t)

Yo[11'(w)cos{wt) + 11"(w)sin(wt)]

(2-60)

11' = (o-o/Yo)sin 8 = G"/w

(2-61)

where:

11"

(o-o/Yo)cos 8

G'/w

(2-62)

64

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The "complex viscosity" is:

7J*(w) == 7J'(W) - i7J"(w)

(2-63)

and:

(2-64)
Since all linear viscoelastic behavior is governed by the
Boltzmann superposition principle, which is based on the single
material function, G(t), it is possible to relate the response to any
sufficiently small or slow deformation to the linear relaxation modulus. In the case of small amplitude oscillatory shear, for example, it
can be shown that G'(w) and G"(w) are the Fourier sine and cosine
transforms, respectively, of the relaxation modulus:

G'(w) =

w /"G(s)sin(ws)

G"(w)

ds

(2-65)

G(s)cos(ws) ds

(2-66)

to
o

If a generalized Maxwell model is used to represent the relaxation


modulus (Equation 2-25), the resulting functions are:

G'(w) =

G"(w) =

G.(WA.)2

i=l

[1 + (wAJ]

G.(wA.)

i=l

[1 + (wAJ2]

E
E

(2-67)

(2-68)

In Section 2.9 we will describe a method for determining the model


parameters [Gi , AJ from experimental complex modulus data.
It is also possible, in principle, to determine the relaxation
spectrum, H(A), from either G'(w) or G"(w) since [1]:

G'(w) =

/X) [H(A)w A2/(1 + w2A2)] d(InA)

(2-69)

G"(w) =

/00 [H(A)wA/(l
-00

(2-70)

-00

+ w2A2)] d (In A)

LINEAR VISCOELASTICITY

65

While the exact inversion of the integrals is not possible, Ferry [1]
has described several empirical methods for determining H(A)
from oscillatory shear data. The major problem here is that G'(w)
and G"(w) can only be determined experimentally over a few
decades of frequency.4 As a result, the reliability of the inferred
H(A) curve decreases rapidly at values of A approaching l/w max
and l/w min , where W max and Wmin are the highest and lowest
frequencies at which experiments have been carried out. Graessley
et al. [12] and Leblans [13] have described elaborate procedures for
obtaining the best possible results from this procedure.
Equations 2-65 and 2-66 allow us to establish the limiting behavior of the storage and loss moduli at high and low frequencies. At
very low frequencies, we find, as expected, that viscous effects
dominate the behavior. In the limit as w approaches zero,
lim G' = lim 71" = 0

(2-71)

lim G" = 0

(2-72)

w--->O

w--->O

w--->O

lim (G"/w) = lim 71' = 710

w--->O

w--->O

(2-73)

where 710 is the "zero shear viscosity" given by Equation 2-19.


Furthermore, from Equation 2-65 it can be shown that the ratio
G'/w 2 approaches a nonzero limiting value, called Ac by Ferry [1],
as the frequency approaches zero. s
(2-74)

By comparison with Equation 2-33 we see that


(2-75)
Since these results are based on the Boltzmann superposition
principle, they are valid for any material exhibiting linear viscoelas4The range can sometimes be extended by use of a technique called time-temperature
superposition, which is described in Section 2.12.
5 For (ws) very small, sin(ws) = ws.

66

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

tic behavior. However, except for linear, monodisperse polymers,


the approach to this limiting behavior is usually very gradual and
may not be obvious.
As the frequency increases, the behavior is more and more
dominated by the elasticity of the melt, and G' increases until it
reaches the glassy modulus, Gg , at frequencies well beyond the
range of most melt rheometers. Other high frequency limits are
given by:
(2-76)
lim r!" = 0
w--->oo

lim G" = lim 71' = 0


(J)~OO

(2-77)

W~OO

In practice, this limiting behavior is not generally observed, because


it is difficult to generate a uniform deformation at sufficiently high
frequencies.
Equations 2-65 and 2-66 show that it is, in principle, possible to
calculate both G'(w) and G"(w) if G(t) is known over a very wide
range of times. This implies that the two functions G'(w) and
G"(w) contain the same information and that it should be possible
to use one to calculate the other. In fact, the relationship is quite
complex, but approximate conversion formulas have been discussed
by Booij and Thoone [10].
2.8.2 Complex Modulus of Typical Molten Polymers

Figure 2-9 is a sketch showing G'(w) curves for resins A, Band C,


whose relaxation modulus curves were shown in Figure 2-3. The
general features of these curves are qualitatively similar to those
seen in the curves of the relaxation modulus (Figure 2-3) where the
short time behavior of the G(t) curve corresponds to the high
frequency portion of the G'(w) curve and vice versa. At the highest
frequencies shown, glassy behavior is exhibited. At somewhat lower
frequencies, molecular rearrangement becomes possible during a
cycle, and there is a transition zone. For the low molecular weight
material (A) we move directly into a terminal zone, whereas for the
high molecular weight material (B), we have a plateau zone. In the
terminal zone, according to Equation 2-65 (or 2-67) the storage
modulus should become proportional to w 2 , while according to
Equation 2-68 the loss modulus becomes proportional to w.

LINEAR VISCOELASTICITY

67

log w

Figure 2-9. G'(w) for three samples of a linear polymer. A is monodisperse with M < M6
B is monodisperse with M Me. and C is polydisperse with Mw Me.

As in the case of the relaxation modulus, we note that in the case


of the material with a broad molecular weight distribution (MWD)6
the plateau modulus is not clearly defined, and there is no sharp
transition to the terminal zone_
Figure 2-10 shows curves of storage and loss moduli for sample
B, which has a very narrow molecular weight distribution and an
average molecular weight far above that at which entanglement
coupling first becomes prominent. For this entangled polymer, G"
passes through a minimum in the plateau zone reflecting the fact
that little dissipation occurs in the region of rubbery behavior. This
also manifests itself in a marked minimum in the curve of ~(w)_
This is in contrast with the behavior of an unentangled melt where
G" is never less than G'_
These examples illustrate the value of small amplitude oscillatory
shear experiments for polymer characterization_
6 See

Appendix B for an explanation of the molecular weight distribution and definitions of


the average molecular weights.

68

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

6~------------------------------------~

2~------~------~------~------~----~
2
6
-2
-4
o
4
log w(rad / s)
Figure 2-10. Typical curves of G'(w) and G"(w) for a narrow distribution linear polymer
(sample B of Figures 2-3,6,9) with M Me.

2.8.3 Quantitative Relationships Between G*(w) and MWD

At the beginning of this Chapter we noted that one of the applications of linear viscoelasticity data is to determine molecular
parameters such as molecular weight (MW) and molecular weight
distribution (MWD). Several empirical relationships between
molecular structure and linear viscoelastic properties have been
proposed for linear polymers [14-19]. Tuminello [16] has reviewed
methods for estimating MWD from viscoelastic data. No such
relationships have been proposed for branched polymers because of
the difficulty of separating the effects of the degree and type of
branching from that of the molecular weight distribution.
Zeichner and Patel [14] found it possible to correlate the breadth
of the MWD for a family of polypropylene resins with the value of
the "crossover modulus," Go which is the value of G' and Gil at the
"crossover frequehcy," We' where G' and Gil are equal. They
defined a "polydispersity index," PI, as follows:
(2-78)

LINEAR VISCOELASTICITY

69

They then found that there was a good correlation between M w/Mn
and PI, with the data falling on a straight line on a double
logarithmic plot.
Wu [15] has proposed a method for determining the MWD from
G'(w) data in the terminal and plateau zones. The basic assumption
is that the cumulative molecular weight distribution curve has the
same shape as the G'(w) curve. Tuminello has reviewed the work
on relations between G'(w) and MWD [16] and has proposed an
improvement of Wu's method which is better suited for use with
polymers having a bimodal MWD [17]. The determination of MWD
from rheological measurements is especially important in the case
of insoluble polymers, as the direct measurement by means of
chromatography is not possible [18, 19].
It is important to note that the various methods that have been
proposed for relating a rheological material function to the molecular weight distribution are empirical, and their validity must be
established experimentally for a given type of polymer.
The calculation of G' and G" for blends of linear polymers is
more straightforward, One empirical blending rule [20,21] has been
tested extensively by Schuch [22].
2.8.4 The Storage and Loss Compliances

As an alternative to the use of the components of the complex


modulus or the complex viscosity to report the results of a sinusoidal shear experiment, the storage and loss compliances, J' and
J", can be used. Here, an oscillatory stress is considered to be the
stimulus, with the sinusoidal strain reflecting the material response.
In fact, the same data that are used to calculate G' and G" can also
be used to calculate J' and J".
(2-79)

(2-80)
We note that 11*1 = 'Yo/ao = I/Gd
Figure 2-11 shows the storage compliance, J', as a function of
frequency for samples A, Band C. The general features of the

70

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

log w

Figure 2-11. Storage compliance versus frequency for three samples of a typical linear
polymer. A is monodisperse with M < Me; B is monodisperse with M Me, and C is
polydisperse with Mw Me.

curves are similar to those of the creep compliance curves with the
progression from low frequency to high corresponding to the progression of the J( t) curves from right to left. An exception is the
behavior of the storage compliance at low frequencies, with J'
approaching a nonzero limiting value equal to the steady state
compliance, J1- Whereas JJ is independent of molecular weight for
the monodisperse samples, it is significantly increased by broadening the molecular weight distribution. The critical value of the
molecular weight at which entanglement effects begin to manifest
themselves in J(t) behavior is given the symbol M~. Plazek et al.
[23] have presented empirical equations for determining J' and J"
from creep data.
2.9 DETERMINATION OF MAXWELL MODEL PARAMETERS

In Chapter 3 we present an empirical model for nonlinear viscoelastic behavior that incorporates the linear relaxation modulus, G(t).
In using this model, it is necessary to represent the modulus by an
explicit mathematical function, and the function most often used is

LINEAR VISCOELASTICITY

71

that of the generalized Maxwell model:


N

G(S)

E Gie-

S/

(2-81)

Aj

i=l

To use such a model, it is necessary to determine a set of N pairs


of values [Gi , Ai] from experimental data.
The experiments most often employed are step strain, in which
G(t) is measured, and oscillatory shear, in which G'(w) and G"(w)
are measured. We consider first the former case. Let us say that
data are available that consist of m sets of values [G k , t k ]. These
data are to be used to determine n pairs of Maxwell parameters
[Gi , AJ The values of the parameters should be chosen so that:
(2-82)
Nonlinear regression is required to determine the parameters
[Gi , A;l, but there is no unique solution [24]. Therefore, these
numbers are only curve fitting parameters and have no physical
significance. They are nevertheless useful for calculating one rheological property from another.
Experimental values of storage or loss modulus can also be used
to determine a set of Maxwell parameters. In this case Laun [25]
suggests that the parameters be chosen such that:
m

[(G'(w k )

G~)2 + (G"(w k )

G\)2]

minimum (2-83)

k=l

where G'(Wk) and G"(Wk) are calculated by means of Equations


2-67 and 2-68. Laun [25] begins by selecting the Ai to have values
equal to integer powers of ten from 10- 4 s to 10 3 s. Then he
determines the G i by use of Equation 2-83. However, when the
resulting parameters are used to calculate various linear material
functions such as G"(w), G(t) and 7J -(t), the resulting curves have
oscillations that do not appear in the corresponding experimental
curves. These result from the arbitrary specification of the Ai
values. The fit can be improved by increasing the number of

72

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Maxwell elements or by using nonlinear regression to determine the


best Ai values [26].
As in the case of Maxwell parameters determined from relaxation modulus data, the parameters determined in this way have no
physical significance and are used only to calculate values of one
rheological material function based on measurements of another.
In Section 2.6 we noted that the determination of the spectrum
function, H(A), from experimental data is an example of an ill-posed
problem, and the inference of a set of Maxwell parameters from
experimental data is another example of an ill-posed problem [27].
This causes significant difficulties in arriving at a meaningful set of
parameters. For example, it is found that the relaxation strengths,
Gi , depend strongly on the choice of relaxation times, Ai. In
attempting to improve the fit by increasing the number of relaxation
times, more and more relaxation strengths are found to be negative,
and the standard error for the fit of the entire curve increases.
Honerkamp and Weese [27] have demonstrated the use of a mathematical technique called the classical Tikhonov regularization to
solve this problem. They show that by using this technique it
is possible to develop a discrete spectrum that converges to the
continuous spectrum as the number of Maxwell elements is increased.
2.10 START-UP AND CESSATION OF STEADY SIMPLE SHEAR AND EXTENSION

While oscillatory shear and creep are the experiments most often
employed to study the linear viscoelastic behavior of molten polymers, there are many others that can, in principle, be used. The
ones described in this section are of interest because they have
been fairly widely used to study nonlinear viscoelastic behavior. In
this context, it is important to know the behavior exhibited in these
experiments in the limiting case of very small strain rate.
In a start-up experiment, a sample initially in an equilibrium state
is subjected to a constant strain rate starting at time t = O. In the
case of simple shear, the data are reported in terms of the "shear
stress growth coefficient" defined as follows:

7]+(t) == u(t)/y

(2-84)

The Boltzmann superposition principle can be used to show how

LINEAR VISCOELASTICITY

73

the shear stress growth coefficient is related to the relaxation


modulus. If we let s = t - t':
11+(t) = {G(s)ds

(2-85)

If the shear stress is suddenly reduced to zero at a time to' during


a start-up experiment, Laun [25] has shown that the ultimate recoil
is given by:

At long times, the stress will become constant, and 11 + will become
equal to 110. By letting t be infinite in Equation 2-85, we can obtain
the relationship between the viscosity and the modulus.
110 =

jooG(s)ds = JOO H(A)d(InA)


o

(2-87)

-00

If the deformation is suddenly stopped at a new time t = 0, after


steady state has been achieved, the stress will decay, and we can
define a shear stress decay coefficient:

(2-88)

From the Boltzmann superposition principle:


(2-89)

For the start-up and cessation of steady simple extension, the


analogous functions are the tensile stress growth coefficient, 11t(t)
and the tensile stress decay coefficient, 11(t), where

< 0]

(2-90)

[i = 0 for t > 0]

(2-91)

[i

0 for

and

74

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

It is also possible to develop relationships between the material

functions of simple shear and those of simple extension. For example, the stress growth and decay coefficients are related as shown in
Equations 2-92 and 2-93.
(2-92)
(2-93)

2.11 MOLECULAR THEORIES: PREDICTION OF LINEAR BEHAVIOR

The material functions defined in this chapter, such as the relaxation modulus, G(t), and the storage modulus, G'(w), are widely
used to compare the rheological properties of two or more polymers. While the Boltzmann superposition principle provides relationships between the different functions, it cannot be used to
predict rheological properties in the absence of experimental data.
There has been some progress in the area of the theoretical
prediction of rheological properties, but the theories that have been
developed to date do not provide a basis for the quantitative
prediction of the properties of polydisperse, high molecular weight
polymers. Nevertheless, these theories are useful in providing guidance regarding general trends in behavior. Furthermore, they provide quantitative predictions of the longest relaxation time, the zero
shear viscosity and the steady state compliance. A brief summary of
the molecular theory of linear viscoelastic behavior is presented in
this section.
2.11.1 The Modified Rouse Model for Unentangled Melts

2. 11. 1. 1 The Rouse Model for Dilute Solutions

Rouse [28] developed a molecular theory for dilute polymer solutions in which a polymer molecule is modeled as a chain of N
straight segments or "submolecules" that act as Hookean springs.

LINEAR VISCOELASTICITY

75

Since the origin of the spring force is Brownian motion, the spring
constant is proportional to the absolute temperature. These segments are connected by "beads" in which the mass of the molecule
is assumed to be concentrated. The motion of the beads through
the solvent gives rise to viscous resistance, which is described in
terms of a friction coefficient, r The combination of the elastic
spring and the viscous resistance gives rise to viscoelastic behavior.
We have seen that for a Maxwell element or a Voigt element,
each of which contains a spring constant and a viscous resistance
parameter, there is only one mode of deformation, and each therefore has a single characteristic time. In the case of the segmented
chain, however, there are many possible modes of deformation
because of the flexibility of the chain, and this gives rise to multiple
relaxation times.
In Rouse's model, there is no "hydrodynamic interaction." This
means that the resistance to bead motion due to the solvent is that
of a particle moving in a liquid in which there are no other
particles. In reality, the flow pattern in the solvent caused by the
motion of one bead has an influence on the resistance of
the solvent to the motion of other beads, even of other beads on the
same chain. Thus, the Rouse theory is not correct even for an
infinitely dilute solution. The theory was modified by Zimm [29] to
account for hydrodynamic interaction. However, the question of
solvent-polymer interaction does not arise in the case of a melt. In
addition to hydrodynamic interaction, another complication that
arises in the case of dilute solutions is the "excluded volume"
effect, which alters the distribution of chain end-to-end distances.
Finally, the Rouse theory does not account for glassy modes of
response, so it is not valid at very high frequencies.

2. 11. 1.2 The Bueche Modification of the Rouse Theory

Bueche [1, p. 225; 30] has suggested that for low molecular weight
molten polymers in which there is no entanglement of molecules,
the Rouse dilute solution theory can be used with rather minor
modification to account for the fact that a polymer molecule is
surrounded not by solvent but by other polymer molecules. This

76

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

modified theory predicts that the relaxation modulus is given by:

G{t)

pRT N
=

L e-

I / Ap

(2-94)

p=l

(2-95)

where:

(= translational friction coefficient per monomer unit

a length characteristic of the chemical structure of the


molecule
degree of polymerization

The viscosity is given by:

770

where:

No
Mo

(2-96)

Avogadro's number
M/P = monomer molecular weight
density

At molecular weights well below those necessary for entanglement to occur, the monomeric friction coefficient, (, is an increasing function of the molecular weight, but there is a significant range
of molecular weights over which it is nearly constant. This implies
that the viscosity increases linearly with molecular weight, and there
is substantial experimental verification of this prediction for low
molecular weight polymers.
By comparing Equations 2-95 and 2-96 we see that the relaxation
times can be related to the viscosity:
(2-97)

LINEAR VISCOELASTICITY

77

In the terminal zone the behavior is governed by the longest


relaxation time (p = 1), and for the Rouse model this is:
(2-98)
This is an important number, as it is a measure of the time required
for internal stresses to relax during annealing. It is also approximately equal to the terminal retardation time, which is the time
required to attain steady flow at constant stress, or the time for
recoil to be accomplished. Note that since Tlo is proportional to M,
as shown by Equation 2-96, the Rouse relaxation times are proportional to M 2
The modified Rouse model predicts the following linear viscoelastic properties:
(2-99)

(2-100)

0.40M
IJ= - - pRT

(2-101)

The last of the above results indicates that the steady state compliance is proportional to the molecular weight for a linear, mono disperse polymer when there are no entanglements (M < M~). This is
in sharp contrast to the behavior observed for highly entangled
(M M~), linear, mono disperse melts, where IJ is found to be
independent of molecular weight.
In the terminal zone, behavior is dominated by the longest
relaxation time. The steady flow properties, Tlo and IJ, whose
Rouse values have already been presented, are examples of properties that are governed by AR. In addition, the following special

78

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

forms of the storage and loss moduli are valid in the terminal zone:

0.4M716W2
1.08PRT] 2
w A2 = - - - M
R
pRT

G'

Gil

1.645pRTwAR/M

w710

(2-102)
(2-103)

Moving our attention to the transition zone, it has been found


[1, p. 189] that if the 2 or 3 longest relaxation times are ignored a
relaxation modulus that approximates the Rouse prediction over a
limited portion of the zone is:

G(t)

Ct- I / 2

(2-104a)

3p RT 71o
27TM

(2-104b)

where:

c=

Since 710 is proportional to M, C is independent of the molecular


weight. Equation 2-104 implies that:
C

H(A) = - A -1/2

,;:;

(2-105)

and

G'( w) = G"( w) =

V7T /2 CW I / 2

(2-106)

These approximate forms are only valid over a limited portion of


the transition zone, because at very short times or high frequencies,
the segmented chain model of a polymer molecule is no longer
valid.
The modified Rouse theory can be extended to polydisperse
systems as long as no species have molecular weights high enough
to participate in entanglements [31,32]. One surprising result is that
the steady state compliance of a blend of two molecular weights can
exceed those of both components, especially when one component
has a much higher molecular weight than the other and this
component has the lower concentration in the blend [33].

LINEAR VISCOELASTICITY

79

2. 11. 1.3 The Bueche -Ferry Law

For high molecular weight polymers, entanglement coupling dominates the rheological behavior, and the Rouse theory is not valid.
However, there is a part of the transition zone in which the
viscoelastic behavior is dominated by configurational changes of
portions of the molecule that are shorter than the distance between
entanglements. These motions are not restricted by the entanglements, and the Rouse theory is thus still valid. The range of times
over which this occurs includes the range over which the approximate spectrum given by Equation 2-105 is valid. Thus, since C is
independent of molecular weight, the curves of G(t), G'(w) and
H(A) are also independent of molecular weight. This can be seen
clearly in Figure 2-9 for the case of G'(w). There is a range of
frequencies covering about one decade in which the storage modulus is independent of molecular weight and in which the curve of
the log( G') versus log( w) is a straight line with a slope of 1/2. Doi
[34] has called this the "Bueche-Ferry law."
2.11.2 Molecular Theories for Entangled Melts

In this section we discuss a type of molecular interaction that only


occurs in high molecular weight melts and concentrated solutions.
This molecular interaction is often described in terms of "entanglements" between molecules or "entanglement coupling." The modern view is that this strong interaction is a purely topological effect
associated with the inability of chains to pass through each other.
However, the term "entanglement" continues to be used, and we
have found it convenient to do so here.
2. 11.2. 1 Evidence for the Existence of Entanglements

The modified Rouse theory has been found to be able to predict


several important features of the behavior of molten linear polymers in which there are no entanglements. One example is the
shape of the curve of storage modulus versus frequency, in particular the transition and terminal zones. As can be seen in Figure 2-9,
however, as molecular weight increases, it reaches a level at which
it begins to have an effect on the curve, shifting the terminal zone

80

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

to lower frequencies and introducing a plateau. This plateau is very


similar to that exhibited by a crosslinked polymer, and it thus
suggests the presence of an interaction between molecules that can
simulate, over a certain range of frequencies, the effect of a rubber
network.
Another manifestation of entanglement coupling is a rather sharp
change in the dependence of viscosity on molecular weight. Whereas
the modified Rouse theory predicts that "70 is proportional to M
(see Equation 2-96), in the neighborhood of a certain molecular
weight, Me> which depends on the chemical structure of the polymer, the viscosity starts to increase much more rapidly with molecular weight, often approaching a proportionality to about the 3.4
power of the molecular weight:

{M>Md

(2-107)

The characteristic molecular weight, Me, is approximately 2Me ,


where Me is the average molecular weight spacing between entanglement points [1, p. 243]. Like the appearance of the plateau, this
change in behavior implies the presence of a new and powerful type
of molecular interaction, one that impedes the flow of one molecule
past another much more strongly than simple hydrodynamic drag.
Entanglement coupling also has a dramatic effect on the compliance, as is shown in Figure 2-6. Whereas the modified Rouse theory
predicts that J~ is proportional to M (see Equation 2-101), for
entangled systems, J~ is found to be independent of M. The
molecular weight at which this change in behavior is observed is
called M~.
2.11.2.2 The Nature of Entanglement Coupling

Ferry [1, p. 243] has summarized the various arguments that have
been put forward to explain the rheological properties attributed to
entanglements. He concludes that these are not due to adherence
between chains at specific points along a molecule but must result
from purely topological constraints on the motion of a molecule. In
other words, entanglement effects are due to the "uncrossability" of
the chains [35]. Figure 2-12 shows one polymer molecule surrounded by its neighboring molecules, the presence of which restricts its motion.

LINEAR VISCOELASTICITY

81

Figure 2-12. Sketch showing a typical molecule (heavy line) together with others that pass
close to it and restrict its motion.

2.11.2.3 Reptation

There has been considerable interest in recent years in molecular


theories of entangled polymers that are based on the concept of
"reptation." This is a mechanism by which the motion of a given
molecule is restricted by the close proximity of segments of many
other molecules.
Using the Rouse segmented chain model to illustrate this concept, we note that in a dilute solution the motion of a bead in a
direction perpendicular to the chain backbone is slowed by the
frictional resistance of the solvent but that motion is otherwise
unimpeded in this direction. However, when a long chain molecule
is surrounded by other long chains it is no longer possible for a
bead to move very far in a direction perpendicular to its chain,
because, unlike a solvent molecule, the segments of the surrounding
molecules cannot simply be pushed aside as each is part of a long
chain. The implication of the uncrossability of chains is that the
only long range motions that are possible are those in which the
chain moves essentially along its own length. The effect of this
topological constraint is similar to that resulting from placing a
chain in a contorted tube [36] that has the same configuration as the
chain. Long range motions are only possible by motion of the chain
along its tube. This motion was called "reptation" by deGennes
[37], after the Latin reptare, to creep, from which the word reptile
also derives.

82

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 2-13. Sketch showing one entire molecule together with the segments of other
molecules that are located near to it and restrict its motion.

2.11.2.4 The Doi-Edwards Theory

A molecular theory of viscoelasticity for molten, high molecular


weight polymers that makes use of the reptation concept has been
developed by Doi and Edwards [38,39]. They start with the Rouse
segmented chain model for a polymer molecule. Because of the
presence of neighboring molecules, there are many places along the
chain where lateral motion is restricted, as shown in Figure 2-13.
To simplify the representation of these restrictions, Doi and
Edwards assume that they are equivalent to placing the molecule of
interest in a "tube," as shown in Figure 2-14. This tube has a
diameter d and a length L. Because the model does not refer to
specific points of entanglement, the molecular weight between

Figure 2-14. Sketch showing the hypothetical tube assumed by Doi and Edwards to be
equivalent in its effect to the segments shown in Figure 2-13.

LINEAR VISCOELASTICITY

83

entanglements, Me' does not appear as a parameter. However,


there is some basis for associating the number of entanglements,
MIMe' with the ratio, Lid.
Doi and Edwards examined the ways in which the chain can
respond to a change in the configuration of its tube caused by a
deformation of the melt. At very short times, the only reaction that
occurs within the chain is the redistribution of extensions among
the segments between the points where topological constraints
(entanglements) are present. The theory predicts that this rapid
relaxation process has a characteristic time, Ae , called the "equilibration time." Once this process is completed the only additional
mechanism available for the molecule to further relieve the stress is
"disengagement," i.e., reptation out of its tube. This is a relatively
slow process with a characteristic time of Ad' which is a diffusion
time. For a long chain, Ad Ae. Another relaxation process that
can, in principle, occur in entangled melts is relaxation of the
contour length, i.e., the retraction of the molecule within its tube.
However, this process makes no significant contribution to linear
viscoelastic properties [39, p. 227].
In this way the theory accounts for the two distinct groups of
relaxation times that are observed for high molecular weight,
monodisperse, linear polymers. For times t < Ae , the theory predicts the same result as is given above for the modified Rouse
theory, in agreement with the Bueche-Ferry law. The model does
not predict a flat plateau, but the decrease in G in the plateau-like
region is only about 20%. Doi and Edwards identify the plateau
modulus with the value of G(t) at t = Ae , i.e.,

(2-108)
With this approximation, for t

Ae the model predicts:


(2-109)

In the plateau and terminal zones the model predicts:


(2-110)

84

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Because of the 1/p2 factor, this result is quite close to a relaxation


with a single relaxation time, i.e., the term for p = 1. This implies
that the relaxation spectrum is quite narrow compared with the
Rouse spectrum.
The longest relaxation time, Ad' is:
a 2 {M3
Ad

M M
e

2 7T 2

kT

(2-111 )

Graessley [40] has derived expressions for other linear properties


predicted by the model:
(2-112)

JO
s

3Me

pRT

5G~

= -- =--

(2-113)

Making use of these expressions, the longest relaxation time, Ad'


can be written in terms of rheological properties:
(2-114)

Comparing these results with experimental observations for linear, monodisperse, entangled polymers, the following general statements can be made. The very strong effect of M on the viscosity is
in qualitative agreement with observation, but the value of 3 for the
exponent is somewhat below the observed value of about 3.4. The
steady state compliance is predicted to be independent of molecular weight for a given polymer. This is in agreement with observation for high molecular weight melts, and it is in sharp contrast with
the prediction of the modified Rouse theory, for unentangled melts,
that JJ is proportional to molecular weight. It is predicted that
JJG~ = 6/5, whereas the observed value is often about 2.
The predicted relaxation modulus at short times and the storage
and loss moduli at high frequencies are not in accord with observation, falling below the experimental curves, especially in the case of

LINEAR VISCOELASTICITY

85

G"(w). These results reflect the very narrow relaxation spectrum

predicted by the theory.


The Doi-Edwards theory is a major step forward from the modified Rouse theory in its ability to predict the most prominent effects
of entanglement in high molecular weight, linear, monodisperse
polymers. However, it is deficient in its ability to make accurate
quantitative predictions of most phenomena. In an effort to preserve the basic tube model while improving quantitative predictions,
several modifications of the theory have been proposed. For example, consideration of an additional relaxation mechanism due to
contour length fluctuations leads to significant improvements in the
shape of the spectrum, the dependency of 710 on m and the value of
Jl [41,42].
A central hypothesis of the Doi-Edwards theory is that the
"tube" retains its identity throughout the disengagement time for
an individual molecule. For a monodisperse system estimates of the
time required for the tube to lose its identity through Brownian
motion (diffusion) of the molecules comprising the tube is much
larger than Ad' and the hypothesis is thus valid. However, in a
polydisperse system this is no longer true, as some of the molecules
making up the tube have a much lower MW than the longest
molecules in the system. In this case, "tube renewal" provides an
additional mechanism for relaxation and thus speeds up the process
[43].
The presence of long chain branching leads to significant theoretical complications. Reptation is no longer possible, and relaxation
occurs primarily through the mechanism of contour length fluctuation [44-46].
2.11.2.5 The Curtiss-Bird Model

In an attempt to avoid some of the assumptions and simplifications


of the Doi-Edwards theory, Curtiss and Bird [47; 48, Chap. 19] have
developed a theory that incorporates not only anisotropic Brownian
motion (reptation) but also anisotropic hydrodynamic drag. Two
empirical parameters are introduced to describe the latter effect: a
"link tension coefficient," E, and a "chain constant exponent," {3.
These are determined by fitting predicted responses to experimental data, with both parameters found often to be between 0.3 and

86

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

0.5. For E = 0, the Doi-Edwards results are predicted. The introduction of f3 brings in a dependence of the hydrodynamic force on
the molecular weight, and this makes it possible to generalize the
model for the case of polydisperse systems. For monodisperse
systems, the theory predicts that the zero shear viscosity is proportional to the longest relaxation time and that both are proportional
to the 3 + f3 power of the molecular weight. Obviously, if f3 is
taken to be 0.4, the often observed dependency of 110 on M is
obtained. The linear steady state compliance for a polydisperse
system is:

(2-115)
Thus, it is strongly dependent on the molecular weight distribution,
in qualitative accordance with experimental observation.
2.11.2.6 Limitations of Reptation Models

The Doi-Edwards model predicts some of the important features of


the behavior of entangled melts, such as the plateau in the modulus
curve and a strong dependence of 110 on M. However, its predictions are not quantitatively correct. The Curtiss-Bird model gives
better fits to experimental data but at the expense of introducing
two empirical parameters.
Kolinski et al. [49] have developed a kinetic model for concentrated polymeric liquids that yields correct scaling relationships, for
example the dependency of the longest relaxation time on M.
Molecular motion simulations based on this model show no preferred motions of the submolecules in the direction of the chain
backbone and thus no evidence of reptation. These controversial
results imply that the tube model is inappropriate for melts, and
further molecular modeling work is needed to provide more conclusive evidence of the validity of the reptation concept.
2.12 TIME-TEMPERATURE SUPERPOSITION

Rheological properties are usually highly temperature dependent.


This means that to obtain a complete picture of the behavior, even
if the behavior is linear, experiments must be carried out at several

LINEAR VISCOELASTICITY

87

temperatures. It is often found that data, for example G'(w) and


G"(w), taken at several temperatures can be brought together on a
single master curve by means of "time-temperature superposition."
This greatly simplifies the description of the effect of temperature.
Furthermore, it makes possible the display on a single curve of
material behavior covering a much broader range of time or frequency than can ever be measured at a single temperature. Materials whose behavior can be displayed in this way are said to be
"thermorheologically simple." We discuss in this section the origins
and implications of thermorheologically simple behavior. At the
end of the section, we examine the limitations of the concept and
explain that it is only useful over certain ranges of time or frequency.
The Rouse theory predicts that temperature affects the relaxation modulus in two ways. First, it changes all the relaxation
times by the same factor (see Equation 2-95). For example, if
A1(T), A2(To)' A3(To)' . .. are relaxation times at a reference
temperature, To, then the effect of changing the temperature to
a different value, T, will be to change these times to
Alar, A2 a r , A3 a r , ... , where a r is a function of T and is equal
to unity at To' Thus:
(2-116)

Note that a r is, in fact, a function of both T and To' The Rouse
theory further indicates that the magnitude of the coefficients, G j ,
are altered as follows by a change of temperature (see Equation
2-94):
(2-117)

Using the above two relationships, the relaxation modulus of a


generalized Maxwell fluid (Equation 2-25) can be rewritten as:

G(t, T)

Tp

oPo

E Gj(To)exp{ -t/[Aj(To)arD

(2-118)

i=l

Or, letting:
(2-119)

88

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

and
(2-120)

we can write:
N

Gr(t r )

L G;(To)exp[ -tr/A;(To)]

(2-121)

i= 1

This equation implies that if Gr is plotted as a function of t" data


taken at various temperatures should all fall on the same curve as
those taken at the reference temperature, To.
In fact, the Rouse theory that leads to the above conclusions
about the temperature dependence of linear properties is not
quantitatively valid for bulk polymers. However, experience has
shown that data for different temperatures can often be superposed
as suggested above. Also, while the aT function cannot be predicted
from first principles it can be determined empirically as a "shift
factor." Thus, if one makes a plot of G (or log G) versus log t, aT is
obtained from the horizontal shift necessary to bring the data for
any temperature T onto the same curve as data for the temperature To.
If Equation 2-121 is assumed to be valid for all values of t r , it can
be used, together with the Boltzmann superposition principle, to
show that all linear viscoelastic properties obey a time-temperature
superposition principle [35]. For example, the viscosity is related to
the relaxation modulus, G(t), according to Equation 2-87, and this
equation can be rewritten as follows in terms of Gr and tr:
(2-122)

But, since:
(2-123)

and
(2-124)

LINEAR VISCOELASTICITY

89

this implies that:


(2-125)
or:
'l70(T)ToPo
'l7o(To)Tp

(2-126)

Thus, the shift factor can be found by measuring the temperature


dependence of the viscosity.
Because the ratio (ToPo/Tp) changes relatively little with temperature in the usual range of melt rheology measurements, an effective shift factor is:
(2-127)

In this case, the shift factor can be determined by measuring the


temperature dependence of the viscosity. An empirical relationship
for 'l7o(T) is the Arrhenius equation:
(2-128)
where Ea is an "activation energy for flow." This equation is often
found to be valid as long as the temperature is at least 100 K above
Tg Closer to the glass transition temperature, the WLF equation
[1, p. 274] has been found useful.
(2-129)
Turning to the storage and loss moduli, Equations 2-65 and 2-66
can be rewritten in terms of reduced variables.
(2-130)
(2-131)

90

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

where G; and G;' have the same relationship to G' and Gil as Gr
has to G. These equations imply that if dynamic data are plotted in
terms of reduced variables, those taken at various temperatures will
all fall on a single master curve, which is the curve of G(w, T). This
procedure is very useful as a means of extending the curve of
G(w, T) well beyond the range of frequencies accessible using
standard laboratory rheometers.
Figure 2-15 shows the storage compliance for poly(n-octyl
methacrylate) [50] at a number of temperatures, while Figure 2-16
is a plot of reduced compliance versus aTw. Figure 2-17 shows the
shift factor as a function of temperature. Note that in the curve of
J;, the data extend from the end of the plateau zone, through the
transition zone and appear to be approaching glassy behavior at the
highest frequencies studied.
In a similar way, time-temperature superposition procedures can
be derived for any type of strain history. If the ratio (ToPo/Tp)
varies little with temperature, superposition can often be accomplished by plotting:
G(t) versus t/a T
G'(w) versus wa T
G"(W) versus wa T
T/'(w)/a T versus wa T
T/"(w)/a T versus wa T
J(t) versus t/a T

We note that for quantities not containing units of time, such as a


modulus, no shift factor is required. This implies that a plot of one
such quantity versus another will be temperature-independent. For
example, plots of G' versus Gil, each point corresponding to a
different frequency, are temperature invariant. Cole and Cole [51]
used a procedure analogous to this in plotting dynamic dielectric
properties, and a plot of G' versus Gil is often called a Cole-Cole
plot. Likewise, a plot of the mechanical loss angle, or of tan 0,
versus G' or Gil is temperature invariant [52]. However, it must be
remembered that no information about frequency dependence can
be obtained directly from such a plot. In other words, no information about the relaxation spectrum can be obtained.

LINEAR VISCOELASTICITY

10- 4

91

r--------------------------------------.

10- 5

10- 9

~----------~------------~----------~
10 2
10 3
10 4

10

FREQUENCY (Hz)-(Iogarithmic scale)

Figure 2-15. Storage compliance data for poly(n-octyl methacrylate) taken at a number of
temperatures. Note that the apparatus used had a useful range of only about two decades of
frequency. All the data are in the transition zone. Data from Ref. 50, which gives curves for
24 temperatures. Figure adapted from Ref. 1, Copyright 1980 by John Wiley & Sons, Inc.
Reprinted by Permission.

92

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

-5

-6

I
Q.

:..,,,
C;
.Q -7

-8

10

11

12

Figure 2-16. Master curve obtained by superposition of the data shown in Fig. 2-15. Note
that the superposed data cover a frequency range of 10 decades! Data from Ref. 50. Figure
adapted from Ref. 1. Copyright 1980 by John Wiley & Sons, Inc. Reprinted by permission.

For crystalline polymers such as polyethylene the exploitation of


time temperature superposition is limited by the need to stay above
the melting point. Furthermore, for highly branched polyethylenes
it has been found by Laun [53] that time-temperature shifting yields
only an approximate superposition of the data, and he concluded
that LDPE is not a thermorheologically simple material. Even in
amorphous polymers there are pitfalls in the measurement of G'
and Gil near Tg [54].
Very careful experiments involving an extended range of time or
frequency have shown that the basic hypothesis of time-temperature superposition, that all relaxation times are equally affected by
temperature, is not entirely valid. Plazek [55] has noted that relaxation times associated with different molecular mechanisms depend
on temperature in different ways. Link and Schwarzl [56] attempted
to apply time-temperature superposition to their data for the creep

LINEAR VISCOELASTICITY

93

8--------------------------------------~

cU

Cl

.2

o
o

25

50

75

100

125

TEMPERATURE,OC

Figure 2-17. Shift factor as a function of temperature for data in Figure 2-15. Data from Ref.
50. Figure adapted from Ref. 1. Copyright 1980 by John Wiley & Sons, Inc. Reprinted by
permission.

compliance of a commercial polystyrene having a broad molecular


weight distribution. The time and temperature ranges used in their
experiments made it possible to measure the response from the
transition zone, through the plateau and well into the terminal
zone. They found that it was not possible to shift the data to
achieve superposition of data in all zones. Figure 2-18 shows what
happened when they determined shift factors by shifting the data in
the terminal zone. Link and Schwarzl [56] then proposed an empirical scheme to permit the superposition of data encompassing all the
zones of behavior.
In spite of its limitations, time-temperature superposition continues to be a useful tool in the analysis of data spanning many
decades of time or frequency. A more thorough discussion of the

94

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

10- 1

~--------------------------------------~

REDUCED TIME

(tIST ), 5

Figure 2-18. Creep compliance versus reduced time for a commercial polystyrene at a
reference temperature of 126.7C. The shift factors have been determined by shifting data in
the terminal zone. Note that the data do not superpose in the transition zone. Adapted from
Ref. 56. Copyright 1987 by Steinkopff Verlag. Reprinted by permission.

temperature dependence of viscoelastic behavior has been presented by Ferry [1, Chap. 11].
In the treatment of experimental data, time-temperature superposition is usually used as a purely empirical procedure. Double
logarithmic plots of data taken at different temperatures are examined to see if shifting can bring the data onto one curve. Then the
shift factor, aT' is determined from the amount of shifting necessary for each curve. It may then be of interest to see if either
Equation 2-128 or 2-129 can be fitted to the aT curve.
Time-temperature superposition has been presented here in the
context of our discussion of linear viscoelasticity. However, it is
often found useful also for the presentation of nonlinear data, for
example, curves of viscosity versus shear rate. This point is examined in Chapters 3 and 4.
2.13 LINEAR BEHAVIOR OF SEVERAL POLYMERS

In order to illustrate the types of linear viscoelastic behavior exhibited by molten polymers, we present here experimental data se-

LINEAR VISCOELASTICITY

95

lected from the large literature on this subject. The largest number
of published results on the effect of molecular weight on rheological
properties are for polystyrene. This is because it is a linear, amorphous polymer and because it is possible to prepare samples having
very narrow molecular weight distributions [57]. For example, Marin
and Graessley [58] have measured some linear viscoelastic properties of nearly mono disperse polystyrene melts.
For the important commercial polymer polyethylene, on the
other hand, it is much more difficult to polymerize samples that
have negligible branching and a very narrow molecular weight
distribution. Thus, most of the published data for polyethylene are
for commercially produced grades.
Figure 2-19 shows the storage modulus data of Onogi et al. [57]
for four, narrow distribution polystyrenes. The values of Mw are as
follows: L14: 28.900; L16: 58,700; L15: 215,000; L19: 513,000, and
the ratio Mw/Mn is less than 1.1 for all four samples. Sample L14
has a molecular weight below that at which entanglement coupling
is present, while the other three samples have significant levels of
7

C?
4
Q.

t;;

..Q

2
L16
0
-6

-5

-4

-3

-2

-1

log( war, s -1)


Figure 2-19. Storage modulus versus reduced frequency for 4 narrow molecular weight
distribution polystyrenes having Mw values of: L14, 2.89 X 10 4 ; L16, 5.87 x 10 4 ; L15,
2.15 X 10 5 ; L19, 5.13 X 10 5 . The reference temperature is 160C. Adapted from Ref. 57.
Copyright 1970 by The American Chemical Society. Reprinted by permission.

96

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

entanglement. Time-temperature superposition was used to present


data over nine decades of reduced frequency. For a number of
narrow distribution polystyrenes with M ;;::: Me, the shift factor was
found to obey Equation 2-129 with:

Cf

C~ =

7.14
112.1

The data do not extend into the glassy zone, because the apparatus
used was not capable of generating sufficiently high frequencies.
At the highest frequencies reached, some degree of molecular
rearrangement was possible during a cycle, and transition zone
behavior is observed. For the unentangled melt, L14, the response
moves directly into a well-defined terminal zone in which G' is
proportional to the square of the frequency, in accordance with
Equation 2-74. Sample L16 has a modest degree of entanglement,
and we see the beginnings of a plateau region, with the terminal
zone shifted to lower frequencies. For samples LIS and L19 there is
a well-defined plateau, where the plateau modulus is independent
of molecular weight. For these narrow distribution samples there is
a sharp transition from the rubbery plateau to a well-defined
terminal region.
In Figure 2-20 the loss modulus is compared with the storage
modulus for sample LIS [57]. In the terminal zone, Gil has a peak
and then falls below the storage modulus over the entire plateau
zone. This decrease in Gil below G' only occurs in entangled melts
and reflects the dominance of elasticity and the relatively small
viscous energy dissipation in the rubbery zone. This phenomenon
can also be seen clearly in plots of mechanical loss angle, or of
tan(i versus frequency, where a well defined minimum occurs in
the plateau region.
The effect of broadening the molecular weight distribution is
illustrated in Figure 2-21, where the storage and loss moduli of a
polystyrene (sample PS7) with Mw = 313,000 and Mw/MN = 1.8
are shown [59]. The most distinctive features of the curves for

LINEAR VISCOELASTICITY

97

8p---------------------------------~

'iii'
D..

iii
::J
3

"C
0

E
.....,
Ol

.Q

o ~~--~--~~~~--~--~~--~--~~
-6

-4

-2

Figure 2-20. Storage and loss moduli for a polystyrene (Ll5) with Mw = 2.15 X 10 5 and
Mw/Mn < 1.1. Adapted from Ref. 57. Copyright 1970 by The American Chemical Society.
Reprinted by permission.

'iii'

D..

iii
::J
3 4
"C
0

5Ol
.Q

3
2
1

-5

-4

-3

-2

-1

log(waT 5 -1)

Figure 2-21. Storage and loss moduli for a polystyrene sample (PS7) having Mw = 3.13 X 10 5
and Mw/Mn = 1.8. Adapted from Ref. 59. Copyright 1970 by The American Chemical
Society. Reprinted by permission.

98

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

7r-------------------------------------~
6

5
PS7

a..

:i

en
.Q

L15

3
2

1~--~----~--~----~--~~--~----~--~
-1
-4
-3
-2
o
3
4
2

log(A, s)

Figure 2-22. Relaxation spectra for polystyrene samples U5 and PS7 as computed by
Masuda et al. from G'(w) and G"(w) data. Adapted from Ref. 59. Copyright 1970 by The
American Chemical Society. Reprinted by permission.

narrow distribution are no longer seen, even though the polydispersity is quite small compared to most commercial polymers. There is
no longer a distinct plateau in the G' curve, and the peak in the Gil
curve has disappeared. Also, there is no longer a sharp transition to
a well-defined terminal zone.
Masuda et al. [59] used an approximate method to determine the
relaxation spectrum from the storage and loss moduli for sample
LIS of Figures 2-19 and 2-20, and for the polydisperse sample (PS7)
of Figure 2-21. These are shown in Figure 2-22. The distinct groups
of relaxation times corresponding to the transition zone and the
beginning of the terminal zone are quite apparent for sample LIS.
This is yet another clear manifestation of entanglement coupling
and the presence of two, well separated, groups of relaxation times.
In the reptation model these correspond to two distinct mechanisms for relaxation; a rapid reorganization of a molecule within its
tube, and the much slower reptation of the molecule out of its tube.

LINEAR VISCOELASTICITY

99

105r-~~--------------------------------~

1 ~----~~--~----~----~----~~--~
3

10-

10

TIME,

t,s

Figure 2-23. Relaxation modulus data of Meissner [61] for LDPE "Melt I." The solid line
was calculated by Laun [62] on the basis of his storage modulus data for the same resin using
a generalized Maxwell model. Adapted from Ref. 62. Copyright 1978 by Steinkopif Verlag.
Reprinted by permission.

The curve for Sample PS7 shows that even a modest amount of
polydispersity results in the loss of the sharp peak in the spectrum
curve. The appearance of a broad flat portion of the H(A) curve for
materials with a broad molecular weight distribution has inspired
the use of an empirical "box" function to fit experimental data [60].
The broadening of the molecular weight distribution also has the
effect of stretching out the spectrum to longer times.
Long chain branching affects the linear viscoelastic behavior in a
way that is qualitatively similar to the broadening of the molecular
weight distribution. In Figure 2-23 we see the relaxation modulus
data of Meissner [61] for a low density polyethylene. In this case we
have both a high degree of long chain branching and a broad
molecular weight distribution. No plateau is evident, and there is
instead a very broad merging of the transition zone into a terminal
zone, with no well-defined transition between zones. This does not

100

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

imply that entanglements are not present but only that there is a
broad spectrum of relaxation times.
The solid curve drawn in Figure 2-23 is based on the calculation
of Laun [62], who used storage and loss modulus data for the same
resin to determine the parameters of the generalized Maxwell
model by use of Equation 2-83. He then used these values to
calculate G(t) by use of Equation 2-25. The oscillations in the
resulting curve result from the arbitrary selection of the Ai values.
The prediction could be improved by using the methods mentioned
in Section 2.8.3.
Marin et al. [63] have presented curves of the spectrum, H(A),
for several linear, monodisperse polystyrenes.

REFERENCES
1. J. Ferry, Viscoelastic Properties of Polymers, Third Edition, John Wiley & Sons,
New York, 1980.
2. N. W. Tschoegl, The Phenomenological Theory of Linear Viscoelasticity: An
Introduction, Springer-Verlag, Berlin, 1989.
3. R. G. Larson, J. Rheol. 29:823 (1985).
4. R. G. Larson, Rheol. Acta 24:327 (1985).
5. M. H. Wagner, Rheol. Acta 15:136 (1976).
6. D. R. Wiff, J. Rheol. 22:589 (1978).
7. C. Y.-c. Lee, D. R. Wiff and V. G. Rogers, J. Macromol. Sci., Phys. B19:211
(1981).
8. K. F. Wissbrun, J. Rheol. 30:1143 (1986).
9. M. Kurata, Macromolecules 17:895 (1984).
10. H. c. Booij and J. H. Thoone, Rheol. Acta 21:15 (1982).
11. H. c. Booij and J. H. M. Palmen, Rheol. Acta 21:376 (1982).
12. W. W. Graessley, W. S. Park and R. L. Crawley, Rheol. Acta 16:291 (1977).
13. P. Leblans, "Constitutive analysis of the nonlinear viscoelasticity of polymer
fluids in various types of flow," Doctoral Thesis, University of Antwerp,
Wilrijk, 1986.
14. G. R. Zeichner and P. D. Patel, Proc. 2nd World Congr. Chern. Eng., Vo!' 6,
p. 373, Montreal, 1981.
15. S. Wu, Polym. Eng. Sci. 25:122 (1985).
16. W. H. Tuminello, "Relating Rheology to Molecular Weight Properties of
Polymers," in Polymer Proc. and Flow Dynamics, Vo!. 9 of Encyc. of Fl. Mech.,
Gulf Pub!., 1989.
17. W. H. Tuminello, Polym. Eng. Sci. 26:1339 (1986).
18. S. Wu, Polym. Eng. Sci. 28:538 (1988).
19. W. H. Tuminello, Polym. Eng. Sci. 29:645 (1989).

LINEAR VISCOELASTICITY

20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.

36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.

49.
50.

101

J. P. Montfort, G. Marin, J. Arman and Ph. Monge, Polymer 19:277 (1978).


J. P. Montfort, G. Marin, J. Arman and Ph. Monge, Rheol. Acta 18:623 (1979).
H. Schuch, Rheol. Acta 27:384 (1988).
D. J. Plazek, N. Raghupathi and S. J. Obron, 1. Rheol. 23:477 (1979).
A. C. Papanastasiou, L. E. Scriven and C. W. Macosko, 1. Rheol. 27:387
(1983).
H. M. Laun, 1. Rheol. 30:459 (1986).
M. Baumgartel and H. H. Winter, SPE Tech. Papers 35:1652 (1989).
J. Honerkamp and J. Weese, "Determination of the relaxation spectrum by a
regularization technique," Macromolecules, submitted 1990.
P. E. Rouse, Jr., 1. Chern. Phys. 21:1272 (1953).
B. H. Zimm, 1. Chern. Phys. 24:269 (1956).
F. Bueche, 1. Chern. Phys. 20:1959 (1952).
G. C. Berry and T. G. Fox, Adv. Polym. Sci. 5:261 (1968).
K Ninomiya, J. D. Ferry and Y. Oyanagi, 1. Phys. Chern. 67:2297 (1963).
H. Leaderman, R. G. Smith and L. C. Williams, 1. Polym. Sci. 36:233 (1959).
M. Doi, 1. Non-Newt. F!. Mech. 23:151 (1987).
W. W. Graessley, "Viscoelasticity and Flow of Polymer Melts and Concentrated Solutions," in Physical Principles of Polymers, Edited by J. E. Mark,
Amer. Chern. Soc., Wash. D.C., 1984.
S. F. Edwards, Proc. Phys. Soc. 92:9 (1967).
P. G. deGennes, 1. Chern. Phys. 55:572 (1971).
M. Doi and S. F. Edwards, 1. Chern. Soc., Faraday Trans. II: 74:1789, 1802,
1818 (1978); 75:38 (1979).
M. Doi and S. F. Edwards, The Theory of Polymer Dynamics, Oxford University
Press, Oxford, 1986.
W. W. Graessley, 1. Polym. Sci., Polym. Phys. Ed. 18:27 (1980).
M. Doi, 1. Polym. Sci., Polym. Phys. Ed. 21:667 (1983); 1. Polym. Sci. 21:667
(1983).
J. Roovers, Polym. 1. 18:153 (1986).
M. Doi, W. W. Graessley, E. Helfand and D. S. Pearson, Macromolecules
20:1900 (1987).
M. Doi and N. Y. Kuzuu, 1. Polym. Sci., Polym. Lett. 18:775 (1980).
D. S. Pearson and E. Helfand, Macromolecules 17:888 (1984).
T. c. B. McLeish, Xth Int. Congr. Rheo!. 2:115 (1988).
C. F. Curtiss and R. B. Bird, 1. Chern. Phys. 74:2016 (1982).
R. B. Bird, C. F. Curtiss, R. C. Armstrong and O. Hassager, Dynamics of
Polymeric Liquids, Vo!' 2, Second Edition, John Wiley & Sons, New York,
1987.
A. Kolinski, J. Skolnick and R. Yaris, 1. Chern. Phys. 86:1567, 7164, 7174
(1987).
W. C. Dannhauser, W. C. Child, Jr. and J. D. Ferry, 1. Colloid Sci. 13:103
(195~.

51. K S. Cole and R. H. Cole, 1. Chern. Phys. 9:341 (1941).


52. C. D. Han and K-W. Lem, Polym. Eng. Rev. 2:135 (1982).

102

53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

H. M. Laun, Prog. Colloid Polym. Sci. 75:111 (1987).


C. L. Rohn, SPE Tech. Papers 35:870 (1989).
D. J. Plazek, Polym. 1. 12:43 (1980).
G. Link and F. R. Schwarzl, Rheol. Acta 26:375 (1987).
S. Onogi, T. Masuda and K. Kitagawa, Macromolecules 3:109 (1970).
G. Marin and W. W. Graessley, Rheol. Acta 16:527 (1977).
T. Masuda, K. Kitagawa, T. Inoue and S. Onogi, Macromolecules 3:116 (1970).
R. D. Andrews and A. V. Tobolsky, 1. Polym. Sci. 6:221 (1951).
J. Meissner, 1. Appl. Polym. Sci. 16:2877 (1972).
H. M. Laun, Rheo!. Acta 17:1 (1978).
G. Marin, J. P. Montfort, J. Arman and Ph. Monge, Rheol. Acta 18:629 (1979).

Chapter 3
Introduction to Nonlinear
Viscoelasticity
3.1 INTRODUCTION

The measurement of linear viscoelastic properties of polymers is a


very useful tool for polymer scientists and plastics engineers. These
properties are readily measured, and they can be related to certain
aspects of the molecular structure of a polymer. Moreover, the
theory of linear viscoelasticity presented in Chapter 2, i.e., the
Boltzmann superposition principle, is useful in providing relationships between the data obtained in different types of experiment.
However, it is important to recall that this theory is only valid
when the deformation is either quite small or very slow. This means
that the deformations used to determine linear viscoelastic properties must be small or slow. In the case of an oscillatory shear
experiment for example, the strain amplitude must usually be less
than about 0.4 in order to determine G'(w) and G"(w). To determine YJo in a steady shear experiment, the maximum shear rate for
linear behavior can be less than 0.01 s - 1 for polymers with longchain branching or a high molecular weight.
For larger or more rapid deformations, the linear theory is no
longer valid, and the response to an imposed deformation depends
on:
1. The size of the deformation.
2. The rate of the deformation.
3. The kinematics of the deformation.

This means that it is not possible to measure a response in one type


of deformation and use the result to predict the response to that
103

104

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

type of deformation (i.e., the same kinematics) unless both the rate
and magnitude of the deformation are the same in both cases. For
example, the relaxation modulus measured for one shear strain
magnitude is not the same as that for a different shear strain
magnitude. Furthermore, since the response depends on the kinematics of the deformation, it is not possible to predict the response
to an extensional deformation on the basis of the results of shear
measurements.
With regard to applications of rheology in the plastics industry,
there are several classes of problem for which the linear theory is
clearly inadequate. First, in plastics forming operations the deformations are generally both large and rapid, and this means that
melt processing behavior often cannot be correlated with linear
viscoelastic properties. Second, it is possible for two polymers of
similar but not identical molecular structure to have linear properties, e.g., curves of G'(w) and G"(w), that are indistinguishable
within experimental error. However, the same two polymers may
differ in their nonlinear properties and thus in their processing
behavior. This means that even if rheological measurements are
only being used for material characterization or quality control, the
linear properties may prove insufficient.
The present preoccupation with linear viscoelastic behavior is
due partly to instrument limitations and partly to the lack of a
suitable theoretical framework to describe and analyze nonlinear
behavior. But we cannot escape the reality that most real world
phenomena are highly nonlinear. The division of viscoelastic behavior into two categories, linear and nonlinear, suggests that nonlinear behavior is somehow exceptional, but this point of view does
not reflect reality. The late mathematician, Stanislaw Ulam, noted
that this is like classifying all animals that are not elephants as
"nonelephants."
Even though there exists at the present time no generally valid
quantitative model for the nonlinear viscoelastic behavior of polymeric liquids, a few basic theoretical concepts have been found
useful in the interpretation of experimental data. The objective of
this chapter is to present these concepts so that they can be used in
the succeeding three chapters on nonlinear melt behavior.
We will begin by examining the various features of the Boltzmann superposition principle to see how they might be modified to

INTRODUCTION TO NONLINEAR VISCOELASTICITY

105

formulate an empirical model for nonlinear viscoelasticity. In Chapter 2, we developed the following form of the principle.
'Tij(t)

G(t - t')Yij(t') dt'

(2-11)

-00

There are three main features of this description of rheological


behavior. The first is the use of the infinitesimal strain tensor, Yij"
This measure of strain is not valid for large deformations. Second,
the relaxation modulus, G(s), is independent of the strain, so that
the response to any part of the strain history is assumed independent of the previously occurring strain history. We do not expect
this to be true for large, rapid deformations. Finally, the effect of
strain history has been accounted for by use of a single integral, and
there is no reason to believe that such a simple procedure will be
adequate to describe the response of a material to large, rapid
deformations.
In this chapter we will look at the consequences of giving up the
first two of these simple features. While this procedure will not lead
us to a general theory of nonlinear behavior, it will provide useful
guidance with regard to the way these nonlinearities first manifest
themselves, and it will suggest procedures for representing experimental data.
3.2 NONLINEAR PHENOMENA

Many rheological phenomena that are of scientific and practical


importance are totally absent from the predictions of the theory of
linear viscoelasticity. Important nonlinear effects manifest themselves even in the simplest flow situations such as steady simple
shear, with linear behavior observed only at very low shear rates.
The predominant nonlinear phenomena are the dependence of the
viscosity on the shear rate and the appearance of a nonzero first
normal stress difference. The latter effect is closely related to the
"Weissenberg effect," which is the tendency of an elastic liquid to
rise up around a rotating rod partially immersed in it.
The richest field of nonlinear phenomena is that of transient
flows involving large strains and large strain rates. Examples of

106

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

nonlinear effects include:


1. the dependence of the relaxation modulus on the strain magnitude,
2. the dependence of 7J +(1) and 7J -(t) on the shear rate, and
3. the dependence of 7J;(t) on the strain rate.
Thus, in the present chapter we must begin to include strain or
strain rate among the independent variables for these material
functions. They then become: G(t, y), 7J +Ct, y), 7J -Ct, y), and

7J;Ct, i).

In the case of large amplitude oscillatory shear, even the definitions of the material functions used to describe linear behavior, e.g.,
C'(W) and C"(w), are no longer useful to describe a material's
response, because these are based on the assumption that the stress
is sinusoidal, and this is not valid for a nonlinear response.
These examples serve to demonstrate that once we enter the
realm of nonlinear behavior, the representation of experimental
data becomes considerably more complex. At the least, an additional parameter, such as strain or strain rate, must be introduced,
and in other cases entirely new material functions must be defined.
Thus, some understanding of the theoretical aspects of nonlinear
behavior is required simply to establish techniques for the meaningful display of experimental results.
3.3 THEORIES OF NONLINEAR BEHAVIOR

The Boltzmann superposition principle provides a basis for the


description of all linear viscoelastic phenomena. Unfortunately, no
such universal theory is available to serve as a basis for the
interpretation of nonlinear phenomena, i.e., to describe flows in
which neither the strain nor the strain rate is small. As a result, we
have no generally valid formulas for calculating values for one
material function on the basis of experimental data for another.
On the other hand, considerable effort has been expended in
improving our understanding of nonlinear behavior in recent years,
and some useful concepts have been developed. While a universal
theory for nonlinear viscoelastic behavior still eludes us, limited
success can be claimed in the development of theories useful for

INTRODUCTION TO NONLINEAR VISCOELASTICITY

107

particular types of deformation. For example, we have achieved


some understanding of behavior in flows that produce only small
departures from linear behavior. Also, semi-empirical formulas for
relating various simple shear material functions have been proposed. These are not spectacular achievements, but they provide
small beacons of light in the otherwise dark world of nonlinear
viscoelasticity.
One approach to the formulation of a nonlinear constitutive
equation is an intuitive one making use of empirical equations for
quantities such as the rates of creation and loss of entanglements. It
should be noted that empirical model building in the area of
nonlinear viscoelasticity is not at all analogous to the devising of
equations for fitting relationships between scalar quantities. Substantial complications arise from two aspects of the problem. These
are the involvement of tensor-valued quantities (stress and strain)
and the fact that the response of the material to a stress or strain
imposed at time t depends not just on these quantities but also on
strains or stresses imposed at previous times.
Because of these complicating factors, it is a significant challenge
simply to establish an acceptable form for a nonlinear constitutive
equation. Certain general hypotheses have proven useful in this
regard. The use of such general criteria to formulate empirical
constitutive equations is called the "continuum mechanics" approach to nonlinear viscoelasticity. Once the general form has been
established, the selection of the specific nature of the equation is
guided by a study of experimental results. Many so-called "continuum" models of nonlinear behavior have been described briefly
in the text by Tanner [1] and in more detail in the comprehensive
monograph by Larson [2].
The alternative approach to the development of nonlinear models is to start from a model for molecular behavior and use statistical mechanics to derive a constitutive equation [3,4]. This approach
is mathematically very complex, and as a result, many simplifying
assumptions are necessary in order to make it possible to obtain an
equation that allows the stress to be calculated from the strain
history. These assumptions result in limitations on the applicability
of the final constitutive equation.
Some of these limitations can be avoided by the use of molecular
dynamics simulations. This involves the use of a supercomputer to

108

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

calculate macroscopic responses directly from the model of molecular behavior without the use of a constitutive equation. This approach is limited by the immense computational power necessary to
model a system of many very long polymer molecules. Also, it yields
only numerical results and not functional dependencies.
In particular, molecular dynamics simulations are not useful in
the modelling of plastics forming operations, although there is a
considerable research literature on the use of constitutive equations
for this purpose. However, the only processes for which such
equations have been successfully used for the quantitative design of
industrial processes are extrusion and injection molding. And in
these cases the rheological model that has been used is a temperature-dependent power law, i.e., an inelastic model.
We have noted above that neither the continuum nor the molecular approach to developing a theory of nonlinear viscoelasticity has
led to a general model for melt flow. Furthermore, the constitutive
equations that have been developed have not proven useful in the
practical modelling of many plastics processing operations.
Nonetheless, the results to date have utility in the following areas:
1. They provide criteria for the appearance of nonlinear effects.
2. They predict the nature of the first departures from linear
behavior.
3. They suggest methods for representing experimental results.
For these reasons, it is important to have some understanding of
certain theoretical concepts, and it is the purpose of this chapter to
introduce these concepts.
3.4 FINITE MEASURES OF STRAIN

First we examine the consequences of replacing the infinitesimal


strain tensor, 'Yij' in the Boltzmann superposition principle, by a
finite measure of strain. The strain tensor defined by Equation 1-37
is not suitable for the description of large deformations and must
be replaced by a finite measure of strain. If 'Yij is replaced by such a
finite strain tensor in the Boltzmann superposition principle, the
result is called a model of "finite linear viscoelasticity" or a "quasi-

INTRODUCTION TO NONLINEAR VISCOELASTICITY

109

linear" viscoelastic model. This is an empirical procedure, but we


will see that it has led to some useful results.
There are, in fact, many possible ways of defining measures of
deformation that can describe finite strains. Some of these can be
ruled out on the basis of continuum mechanics concepts. Rational
mechanics is the branch of continuum mechanics that examines
possible forms for constitutive relationships. Starting from general
hypotheses about material behavior, implications regarding the
mathematical representation of constitutive equations are derived.
The most important such hypothesis is the principle of material
indifference, which states that a material's rheological behavior
reflects a basic physical property and therefore cannot depend on
the frame of reference used to describe the behavior. 1

3.4.1 The Cauchy Tensor and the Finger Tensor

The principle of material indifference can be used to test finite


measures of strain to see if they are acceptable candidates for use
in formulating a theory of finite linear viscoelasticity. Two measures
of strain that satisfy this criterion and that have been found
particularly useful in polymer rheology are the Cauchy tensor,2
Cij(t p t 2 ), and the Finger tensor,3 Bij(t 1 , t 2 ). The time arguments
have the following significance: t 1 is the time at which a material
element is in its reference configuration, and t2 is the time at which
the strain is evaluated, relative to the configuration at time t}. The
mathematical definitions of these tensors and a discussion of their
physical significance are presented in Appendix A.
For most of our purposes in this book it will be sufficient to know
the components of these two tensors for the two types of deformation most used for rheological measurements: simple shear and
simple extension. For simple shear, the components of the Cauchy

t This assumes the material to be isotropic in its rest state, which is generally true for
homogeneous melts but not for reinforced materials or liquid crystals.
2This quantity is sometimes called the Cauchy-Green tensor.
3 The Finger tensor is the inverse of the Cauchy tensor and can also be written as Ci";t<tt, t 2 ).
The symbol, Bij' is used by some authors to refer to the Green tensor, which is not used in
the present book.

110

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

and Finger tensors are shown below.

[1'(t 2) - l'(t 1)]

{I + [1'(t 2) -

l'(t1)]2}

o
(3-1)
and

(3-2)
For simple extension, the components of the Cauchy and Finger
tensors are:

Cij(t 1, t 2) =

[e''''')-'''')]
0
0

_[e""')-~")]
0

B i /t 1, t 2) -

0
e-[E(t 2 )-E(tl)]

e-""L,)] ]

(3-3)

e-"',L,,)]]

(3-4)

0
e-[E(tl)-E(t2)]

3.4.2 Strain Tensors

We note from Equations 3-1 to 3-4 that the components of the


Cauchy and Finger tensors are not zero when a material is in its
undeformed state. In fact, in the undeformed state, both tensors

INTRODUCTION TO NONLINEAR VISCOELASTICITY

111

become equal to the unit tensor, whose components are shown


below:

[~

o
1

~l

For this reason, it is sometimes convenient to use the Cauchy strain


tensor and the Finger strain tensor, defined as follows:

where

()ij

Cauchy strain tensor ==

C ij -

()ij

(3-5)

Finger strain tensor ==

()ij -

Bij

(3-6)

is the "Kronecker delta" defined as follows:

(i = j)

(3-7a)

(i

(3-7b)

=1=

j)

By reference to Equations 1-44 (letting y == y(t 2) - y(tl))' 3-1,


3-2, 3-5 and 3-6, we see that for infinitesimal simple shear deformations:
(3-8)
Furthermore, for infinitesimal simple extensional deformations, we
note from Equations 3-3 and 3-4 that the nonzero components of
C ij and Bij are exponentials. From the first term of the series
expansion of eX, which is 1 + x, we see that for very small values
of E:

e 2e = 1 + 2E

(3-9a)
(3-9b)

By comparison with Equation 1-47, we see that Equation 3-8 is also


valid for very small simple extensional flows. It can in fact be shown
that it is valid for all infinitesimal deformations, and this justifies

112

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the use of 'Yij in the Boltzmann superposition principle as the


correct measure of strain for linear viscoelastic behavior.
3.4.3 Reference Configurations

In the above introduction of relative strain tensors, the meaning of


the two time arguments, t1 and t2 is as follows: BJt 1, t 2) is the
Finger strain at time t2 relative to the configuration of the fluid at
time t 1. Thus, time t 1 defines the reference state with respect to
which the strain is measured. In an experiment in which a sample is
initially at rest for a long period of time, so that it is free of
deforming stresses, it is convenient to set t1 equal to the time, to' at
which deformation begins.
This is a suitable choice if one wishes only to describe a particular experiment using a component of the Cauchy or Finger tensor
as the independent variable. For example, it is sometimes of interest to plot the shear stress growth coefficient 7J +, as a function of
shear strain, 'Y, rather than of time, where 'Y =
It is also
adequate for a calculation of stress within the theory of linear
viscoelasticity. This is because this theory is only valid for very small
deformations, so the configuration of a fluid element never deviates
significantly from that at the beginning of the deformation.
However, for large deformations the use of the initial state as the
reference state only has absolute significance at the very beginning
of the imposed deformation, i.e., at times just longer than to' For
large deformations occurring over a long period of time, the configuration at time to becomes less and less relevant due to the
fluid's fading memory, and eventually loses its relevance altogether.
Our present objective is to generalize the theory of linear viscoelasticity so that it can describe the response of a molten polymer
to large, rapid deformations, i.e. we wish to know how the stress at
time t, TJt), depends on deformations occurring at past times, t',
where t' ~ t. Here the only configuration that has any unique
significance is that which exists at the time t. This suggests that we
should measure all strains relative to the configuration of the fluid
element at time t. Therefore, in using one of the relative strain
tensors to formulate a constitutive equation, we take t2 to be t', the
time for which the contribution of the strain to the stress is being
evaluated, and t 1 we take to be t, the time at which the stress is to

rt.

INTRODUCTION TO NONLINEAR VISCOELASTICITY

113

be determined. In other words, the fluid element is in its reference


configuration at time t. The Finger tensor, evaluated at time t', is
thus Bij(t, t').
3.4.4 Scalar Invariants of the Finger Tensor

A vector has one "scalar invariant," i.e., a scalar that can be


determined from the components of a vector, but whose value,
unlike those of the components of the vector, is independent of the
coordinate system used to describe the vector. The scalar invariant
of a vector v is its magnitude, v, given by:

v = Vv; + vi + vi

(3-10)

A second order Cartesian tensor has three scalar invariants. In this


book, we will make use of the scalar invariants of the Finger tensor
for an incompressible fluid. For a given deformation, these can be
calculated as follows:

(3-11)
(3-12)

(3-13)
Since B jj is a function of the two times t and t', the scalar
invariants also depend on these two quantities.
We will make quantitative use of the invariants only in connection with simple shear and simple extension. Therefore, we present
here the values of 11(Bjj ) and liBj) for these deformations. For
simple shear, with shear strain 'Y, we have, using the diagonal
components from Equations 3-1 and 3-2:
(3-14)
For simple extension, we can use Equations 3-3 and 3-4 to show

114

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

20

-'"
10

NO DEFORMATION

10

20

30

Figure 3-1. Relationship between [1(B i ) and [2(Bij) for various types of deformation. All
possible deformations of a constant density liquid are within the shaded areas.

that:
(3-15)
(3-16)
The invariants of the Finger tensor can be used to classify the
kinematics of various types of deformation [5]. This is shown in
Figure 3-1, where every possible deformation of a constant density
fluid corresponds to some point in the shaded regions.
3.5 THE RUBBERLIKE LIQUID

Either the Cauchy or the Finger tensor, or a combination of the


two, can be used to formulate a theory of finite linear viscoelasticity, and rational mechanics provides no guidance as to which choice
will yield the best model of reality. Experimental observation,
however, has shown that if only one tensor is to be used, the Finger

INTRODUCTION TO NONLINEAR VISCOELASTICITY

115

tensor gives the best indication of behavior at least for deformations just beyond the region of linear viscoelasticity.
3.5.1 A Theory of Finite Linear Viscoelasticity

If the Finger tensor is used to generalize the Boltzmann superposi-

tion principle and thus formulate a theory of nonlinear viscoelasticity, the result is as follows 4 :
7

i/t) =

m{t - t')Bij{t, t') dt'

(3-17)

-00

Lodge called Equation 3-17 the "rubberlike liquid" model and has
reported the predictions of the model for many types of deformation [6].
The relation between the memory function and the relaxation
modulus of the rubberlike liquid can be found by using Equation
3-17 to calculate the shear stress resulting from a step shear strain
of magnitude y, imposed on a rubberlike liquid at time t = O. From
Equation 3-2, we see that for any simple shear deformation B 21 (t, t')
is [y(t) - y(l')]. The shear stress, (J" (equal to 721) is thus given by:

(J"{t) =

m{t - t')[ y{t) - y{t')] dt'

(3-18)

-00

For the specified deformation we have, for

[y{t) - y{t')] =

[y{t) - y{t')] = 0

> 0:

<0

(3-19a)

t' > 0

(3-19b)

t'

The shear strain for t' > 0 is zero, because no deformation occurs
between the times t' and t. Thus:

(J"{t) =

fO

m{t - t') dt'

(3-20)

-00

4The material is in its reference configuration at the time t, at which the stress is being
evaluated.

116

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

or, using the definition of the relaxation modulus:

a(t)

G(t) == -

or, letting s

'Y

m(t - t')dt'

(3-21)

-00

t - t':

(3-22)

One can show that:


m(t - t') =

dG(t - t')
dt

(3-23)

or in terms of s:
m(s) =

dG(s)

(3-24)

ds

We note that the relaxation modulus of the rubberlike liquid is


independent of the strain. This means it is the same modulus as
that which governs the relaxation after a strain sufficiently small
that the Boltzmann superposition principle is valid. In other words,
G(t) in Equations 3-21 to 3-24 is the relaxation modulus of linear
viscoelasticity. This means that the memory function, met - t'), is a
linear viscoelastic property. Thus, it can be written in terms of the
continuous spectrum, H(A):
m(s) =

H(A)
f __
eA
00

s/ A

dOn A)

(3-25)

-00

The rubberlike liquid (Equation 3-17) is not a satisfactory general


model of nonlinear viscoelastic behavior, but it is important because
it predicts the main features of the first deviations from linear
behavior. In addition, several more complex constitutive equations
that have been proposed for polymeric liquids are of this general
form. Thus, the rubberlike liquid provides a useful point of departure for our introduction to nonlinear viscoelasticity.

INTRODUCTION TO NONLINEAR VISCOELASTICITY

117

3.5.2 Lodge's Network Theory and the Convected Maxwell Model

Using hypotheses and mathematical techniques that had proven


useful in the theory of rubber elasticity, Lodge [7] derived a constitutive equation that is a special case of the rubberlike liquid
equation. He assumed that the strong interactions between polymer
molecules in a melt could be looked upon as forming a network in
which entanglements act as temporary crosslinks. This made it
possible to use mathematical developments originally formulated to
describe rubber elasticity. In order to introduce a dependence on
time, which is not present in the theory of rubber elasticity, Lodge
assumed that network junctions are continuously being created and
lost. In order to avoid the introduction of empirical constants, he
further assumed that the loss and reformation of temporary
crosslinks are thermal phenomena unaffected by the deformation of
the material.
The memory function that arises from Lodge's network theory is:

m(t - t')

N G.
- t') 1
E
-' exp [(t
---

i=l Ai

(3-26)

Ai

By use of Equation 3-21 it can be shown that the relaxation


modulus corresponding to this memory function is identical to that
of the generalized Maxwell model. The constitutive equation that
results from Lodge's network theory isS:
(3-27)
Lodge's network model looks upon the melt as a collection of
network strands rather than a collection of molecules. Therefore,
molecular parameters such as chain length and friction coefficient
do not appear, and the theory does not predict the relaxation
spectrum. Neither does it predict the effects of molecular structure
on rheological properties. However, because of its simplicity and
5Stress fields calculated by the use of Equation 3-27 for specified kinematics are also
solutions of the "upper convected Maxwell model," which is a finite strain generalization of
the differential equation form of the generalized Maxwell model described in Section 2.6.1.

118

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the absence of empirical parameters, Lodge's network model has


stimulated considerable interesting discussion on melt rheology.
3.5.3 Behavior of the Rubberlike Liquid in Simple Shear Flows

We will now examine the behavior of the rubberlike liquid defined


by Equation 3-17 for simple shear flows. To do this, we can use the
components of the Finger tensor shown in Section 3.4.1. First, we
consider the shear stress, (T(t), in a simple shear flow. From Equations 3-17 and 3-2, we have, for any simple shear deformation:

(T(t) == 7 21 (t)

met - t')[ yet) - yet')] dt'

(3-28)

-00

Expanding this, integrating the second term by parts, and noting


that G( 00) = 0, one can show that this is equivalent to

(T(t)

G(t - t') dy(t')

(3-29)

-00

Thus, the predictions of the rubberlike liquid for the shear stress in
any simple shear flow are the same as those given by the theory of
linear viscoelasticity.
Turning now to the normal stress differences in simple shear, by
inserting the appropriate components from Equation 3-2 into Equation 3-17 we obtain:

N}(t)

met - t')[ yet) - y(t,)]2 dt'

(3-30)

-00

(3-31)

We see that the first normal stress difference, N}(t), in any simple
shear deformation is positive, which is a nonlinear effect not predicted by the Boltzmann superposition principle. However, the
second normal stress difference, N 2 , is predicted to be zero. While
experimental observations do show that N} is positive, they also
show that N2 is not zero and has a negative value somewhat smaller
in magnitude than N 1

INTRODUCTION TO NONLINEAR VISCOELASTICITY

119

3.5.3.1 Rubberlike Liquid in Step Shear Strain

For a step strain of magnitude 'Y, the shear stress for a rubberlike
liquid is identical to that of linear viscoelasticity:

(T(t)

G(th

(3-32)

From Equations 3-19, 3-30 and 3-31, we have, for the first and
second normal stress relaxation functions;
(3-33)

NzCt, 'Y)

(3-34)

The result for Nl is clearly a departure from linear viscoelastic


behavior.
By combining Equations 3-32 and 3-33 we obtain:

(3-35)

Equation 3-35 is called the Lodge-Meissner relation [8,9]. It has


often been observed to hold for molten polymers well beyond the
normal regime of validity of the rubberlike liquid equation.
3.5.3.2 Rubberlike Liquid in Steady Simple Shear

Start-up of steady simple shear is a homogeneous deformation in


which a fluid at rest is suddenly subjected to steady simple shear at
t = O. For the shear stress growth function the prediction of the
rubberlike liquid equation is the same as that given by the Boltzmann superposition principle:

(T+(t)

ytG(t) + Y{m(s)sds
o

y{G(s) ds
0

(3-36)

120

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

For the first normal stress growth function we have:

N((t) = y 2t 2G(t) + y2[m(s)s2ds = 2 y 2[G(s)sds (3-37)


o
0
while N;(t) is zero. If the shear rate remains constant until a+(t)
and N(t) become independent of time, their steady state values
are, respectively, the product, YT}, and the first normal stress
function. The viscosity prediction is the same as that given by the
Boltzmann superposition principle.
For the first normal stress function we obtain, from Equation
3-37:
(3-38)
The prediction for N 1( y) indicates that the first normal stress
difference is proportional to the square of the shear rate. This
implies that the first normal stress coefficient is independent of
shear rate:
\(1'1

N1
== --=-2
y

foom(s)s2ds
0

2 fOO G(s)sds
0

(3-39)

It has been observed that in the limit of very small shear rates, N1

for polymer melts usually does become proportional to y2, and it is


this observation that originally inspired the definitions of the normal stress coefficients. Thus, Equation 3-39 gives the correct limiting low shear rate behavior of molten polymers and thus provides
an accurate prediction of \(1'1,0.

This result can be expressed in terms of the relaxation spectrum


function as follows:
\(1'1,0

= 2fooH(it)itdit = 2foo H(it)it 2 d(Init)


o

-00

(3-41)

INTRODUCTION TO NONLINEAR VISCOELASTICITY

121

By comparison with Equation 2-87 we note that the first normal


stress coefficient is related to a higher moment of the relaxation
spectrum than the viscosity. This fact will prove useful in Chapter 4
when we wish to relate these properties to the molecular weight
distribution. Making use of Equation 2-33, the integral in Equation
3-40 can be rewritten in a form involving the linear viscoelastic
value of the steady state compliance:
(3-42)

Furthermore, by making use of Equation 2-65, and noting that


sin(s) = s when s is very small, Equation 3-40 can also be written
in terms of the limiting low frequency behavior of the storage
modulus:
'1'1,0

= 2 lim

w--->o

[G~]
W

(3-43)

Thus, from Equations 3-42 and 3-43 we see that the rubberlike
liquid model gives an accurate prediction of the low-shear rate
limiting value of a nonlinear viscoelastic property, '1'1 0' in terms of
linear viscoelastic properties.
'
For "cessation of steady shear," the shear rate is suddenly
reduced to zero after steady stresses have been established in
steady shear flow. The relevant material functions are:

1 G(s)ds

(3-44)

'l'l-(t) = fO met - t')(t,)2 dt'

(3-45)

71-(1)=

00

-00

These predictions are valid for actual melts in the limit of very
small shear rates.
3.5.3.3 Rubberlike Liquid in Oscillatory Shear

For a sinusoidal shear strain, the shear stress is the same as in the
case of linear viscoelasticity, as is predicted by Equation 3-29.
However, there now appears a first normal stress difference that is

122

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

sinusoidal, with a frequency of 2w and an average value over a


cycle of y;G"(w) [6].
3.5.3.4 Constrained Recoil of Rubberlike Liquid

In Section 2.7 constrained recoil following simple shear deformation was described, and the predictions of the theory of linear
viscoelasticity for the ultimate recoil or "recoverable shear," Yoo'
were presented. Since these results involve only the shear stress,
they are also valid for the rubberlike liquid. However, Nl is not
equal to zero in the rubberlike liquid. This means that in order to
carry out a constrained recoil experiment, it is necessary to maintain a nonzero value of N 1 Furthermore, Laun [10] has shown that
it is possible to relate the ultimate recoil to the first normal stress
functions given in the previous section. For start-up of steady
simple shear,
(3-46)

where to is the time during the start-up experiment when the shear
stress is suddenly reduced to zero, leading to recoil.
If we let to approach 00, we obtain the ultimate recoil for steady
simple shear:
(3-47)

We expect this equation to be valid for actual materials in the limit


of very small shear rates. Thus, an equation expected to have
general validity is:
(3-48)

3.5.3.5 The Stress Ratio (N 1/u) and the Recoverable Shear

The ratio of the first normal stress difference to the shear stress at a
particular shear rate is sometimes used as a measure of the "elastic-

INTRODUCTION TO NONLINEAR VISCOELASTICITY

123

ity" of a melt at that shear rate. This concept has its origin in the
classical theory of rubber elasticity. For a purely elastic, linear
rubber, the shear stress resulting from a simple strain of magnitude
'Y is:
(1' = G'Y

(3-49)

To generate this deformation it is also necessary to impose a


nonzero first normal stress difference having a magnitude of
(3-50)

For a crosslinked material, all of the strain is recoverable on release


of the deforming stresses. Therefore, the recoverable shear (ultimate recoil) is equal to 'Y, and from the above two equations this is:
NI
'Yoo = 'Y = -(1'

(3-51)

For the rubberlike liquid, the recoverable shear for steady simple
shear is given by Equation 3-47, which can be written as:
NI
'Yoo = 2(1'

(3-52)

We note that this is only half the recoverable shear for an ideal
rubber for the same value of NIl (1'. This is because disentanglement and reentanglement occur during the recoil process so that
parts of the strained network strands are continuously replaced by
unstrained network strands. The quantity on the right hand side of
Equation 3-52 is sometimes called the "recoverable shear," but as is
shown in Chapter 5 this quantity is not equal to the actual recoverable shear except at very low shear rates.
3.5.4 The Rubberlike Liquid in Simple Extension

For step shear in extension, the tensile relaxation modulus is given


by:

(t,l3) =

2e
(e e-e) G(t)
13

(3-53)

124

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

From Equation 3-9 it can be seen that in the limit of very small
Hencky strains, this reduces to the linear viscoelastic behavior
described by Equation 2-15.
In practice, step strain in extension is not a practical experiment
for a molten plastic, and the test that has been most used is tensile
start-up flow. For this flow, we can use the appropriate components
of the Finger tensor, as given in Equation 3-4 together with Equation 3-17 to obtain the tensile stress growth function for a rubberlike liquid. The result is as follows:
(3-54)
This is clearly different from the linear viscoelastic result given as
Equation 2-92.
To obtain a clearer picture of the behavior of the rubberlike
liquid at the start-up of extensional flow, we can use the relaxation
modulus of the Maxwell fluid, as given by Equation 2-25 to obtain:
Tl;(t,i)

(1 ~~iA )[1-

e-(l-2iA)I/A]

+( 1: iA )[1-

e-(1+iA)I/A]

(3-55)

where Tl is equal to GoA for a Maxwell fluid.


Figure 3-2 is a plot of (Tl+/Tl) versus (t/A) for various values of
(iA). The following interesting features can be noted. First, at very
low dimensionless strain rates (iA), we recover the result given by
Equation 2-92 for linear viscoelastic behavior. Second, the curves
for nonzero iA rise above the linear curve and reach steady state
values as long as i < 1/2A. The steady state value of Tl +(t, i) is
the "extensional viscosity". From Equation 3-55 this is:
TlE(i)

3Tl
(1 - 2iA)(1

+ iA)

(3-56)

The fact that no steady state stress is reached when i ~ 1/2A is


reflected here in the prediction of an infinite or negative value of

INTRODUCTION TO NONLINEAR VISCOELASTICITY

125

10r--------------------------r-------r----------~

0.8

1.2

1.6

2.4

2.8

tlA
Figure 3-2. Tensile stress growth function divided by the viscosity for a rubberlike liquid with
a single-exponential relaxation modulus. The lower envelope curve corresponds to linear
viscoelastic behavior, which is predicted by the rubberlike liquid model as a limiting case
when eA -> O.

TJE(i) for these cases. These results have no physical significance,

however, as unless the stress has a steady state limiting value, the
extensional viscosity is inappropriate for describing the behavior of
the material.
For a relaxation modulus expressed in terms of a discrete spectrum of relaxation times, the tensile stress growth function is given
by Equation 3-55 with the right hand side replaced by a sum of
terms, each one like the right hand side but with TJ replaced by TJj
and A replaced by Aj In this case, u;(t, i) and TJ;(t, i) increase
with time without bound when i exceeds the reciprocal of twice the
longest relaxation time.
It has been observed (see Chapter 6) that high molecular weight
polymers have tensile stress growth functions that begin their deviation from linear behavior by rising above the linear viscoelastic

126

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

curve, Equation 2-92, in the manner predicted by the rubberlike


liquid model. In the case of LDPE, a highly branched polymer, the
stress increases especially markedly, in general agreement with
the rubberlike liquid behavior. But in all cases observed to date, the
measured stress has a limiting value, so that the unbounded increase predicted by the model when i is above some critical strain
rate is not observed.
3.5.5 Comments on the Rubberlike Liquid Model

The rubberlike liquid model does not provide a quantitative description of the behavior of molten polymers undergoing large,
rapid deformations. In particular we note the following as examples
of deficiencies in its predictions.
The viscosity is independent of shear rate.
The first normal stress coefficient is independent of shear rate.
The second normal stress function is zero at all shear rates.
The tensile stress growth function increases without limit
when the strain rate exceeds the reciprocal of twice the
longest relaxation time.
5. The tensile stress growth function always lies above the linear
viscoelastic curve at nonzero values of i.
1.
2.
3.
4.

These predictions are clearly not in accord with experimental


observations. However, the rubberlike liquid model is important for
the following reasons. First, it is a simple equation having no
parameters that must be determined by observing nonlinear behavior. In addition, it gives the correct low shear rate limiting dependence of the first normal stress function on shear rate and thus
provides relationships between '1'1 0 and the material functions of
linear viscoelasticity. Finally, it provides a basis of comparison for
describing the nonlinear viscoelastic behavior of real materials.
Specifically, the actual behavior can be compared with the predictions of the model and the deviation then used as a nonlinear
characterizing function.
For example, Wagner [11] has suggested that in evaluating the
behavior of a melt at the start-up of steady simple extension, the
tensile stress growth function, 7JI(t, i), should be compared with

INTRODUCTION TO NONLINEAR VISCOELASTICITY

127

the behavior of a rubberlike liquid having the same memory function, rather than with the predictions of linear viscoelasticity. He
further proposes that when the data fall below the curve for the
rubberlike liquid, the behavior should be described as "strain
softening" even if the data lie above the linear viscoelasticity curve.
This point of view stems from Wagner's interpretation of deviations
from rubberlike liquid behavior in terms of a strain-induced reduction in the entanglement density.
3.6 THE BKZ EQUATION

Lodge's rubberlike liquid model is the simplest theory of nonlinear


viscoelasticity that is capable of predicting most features of the first
appearance of nonlinear behavior, when both the size and the rate
of the deformation exceed the ranges in which linear behavior is
observed. As was mentioned in Section 3.3, all attempts to develop
more general nonlinear continuum models have led to much more
complicated constitutive equations, and none that have been proposed to date have been found to have any degree of universality in
their predictive ability.
However, there is one continuum model that has proven especially important in the development of present day ideas about
nonlinear viscoelasticity, and we wish to say a few words about it.
This is the BKZ equation proposed by Bernstein, Kearsley and
Zapas [12].6 Making use of concepts originally used in the development of the theory of rubber viscoelasticity, they proposed the
following form for the constitutive equation of a viscoelastic
material.
Tij =

-00

au ij (t, t') - 2au- Bij(t, t') dt'


2-C
all
a12

(3-57)

where u is a time-dependent elastic energy potential function:

u =

U(/l'

12 , t - t')

(3-58)

6 The same concept was independently developed by A. Kaye and discussed in Note No. 134
of the College of Aeronautics, Cranford, England in 1962.

128

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

This function must be determined experimentally, by the study of


large, rapid deformations. For this reason, the basic form of the
model shown has been little used.
A more useful special case of the BKZ equation was formulated
to incorporate the observation that stress relaxation data for
crosslinked rubbers could often be described by a relaxation modulus that is a product of a time-dependent term and a strain-dependent term. This behavior can be incorporated into a BKZ type
equation by the introduction of a time-independent potential function as follows:

(3-59)
This leads to the "factorable BKZ model":
Tij =

au (t, t') - 2-Bjit,


au
met - t') [ 2-Cjj
t') dt' (3-60)
all
aI2

-00

The potential function must again be determined by experiment.


We note that the Lodge rubberlike liquid model can be looked
upon as a special case of Equation 3-60. Tanner [13] has written an
interesting review of the extensive literature inspired by the BKZ
theory.
Most tests of the BKZ equation involve measurements of the
shear stress and the first normal stress difference in simple shear
flows. For these quantities, the predictions of the factorable BKZ
equation are the same as those of Wagner's equation, which is
discussed in the next section.
3.7 WAGNER'S EQUATION AND THE DAMPING FUNCTION
3.7.1 Strain Dependent Memory Function

Many continuum models have been formulated by modifying the


rubberlike liquid model to improve its ability to fit nonlinear
response data. The generalizations that have been most explored
are those in which the simple, single integral form is preserved but
in which the memory function is permitted to depend on various
variables describing the deformation process. Among the variables

INTRODUCTION TO NONUNEAR VISCOELASTICITY

129

that have been proposed are:


1. The history of the strain rate in the interval t' to t.
2. The history of the strain in the interval from t' to t.
3. The history of the stress in the interval from t' to t.
4. The stored elastic energy at time t'.

The approach that has been found to be most useful is to let the
memory function depend on the strain as well as on time. Since the
memory function is a scalar quantity, while the strain is a tensorvalued quantity, this can only be accomplished by letting the memory function depend on the scalar invariants of the Finger tensor,
which are defined by Equations 3-11 and 3-12.
Thus, the memory function is assumed to take the following
form:
(3-61)
and the constitutive equation is:

Tiit) =

M[(t - t'), Il(BiJ, 12 (Bij)] Bij(t, t') dt' (3-62)

-00

By comparison with Equation 3-57 we see that this is a special


case of the BKZ model in which the term involving the Cauchy
tensor has been omitted. This greatly simplifies the use of the BKZ
model, but because the memory function is no longer derivable
from an energy potential, the thermodynamic consistency that is
built into the BKZ equation is lost. This makes it possible for
Equation 3-62 to predict responses that violate the second law of
thermodynamics in the case of very rapid cyclic deformations.
Furthermore, Equation 3-62 predicts that the second normal stress
difference in any simple shear flow is zero, whereas experimental
data for steady simple shear flows of a number of materials indicate
that N2 is negative and has a magnitude about 10 to 30% that
of N l
Because the second normal stress difference is thought to reflect
an aspect of melt behavior that is not very important in many flows
of practical importance, and because the inclusion of the Cij term
makes the equation more difficult to use, Equation 3-62 is the most
used special case of the BKZ equation. The omission of the Cij

130

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

term also means that Equation 3-62 can be looked upon as a


generalization of the rubberlike liquid equation.
It was noted many years ago that stress relaxation data for
crosslinked rubbers could often be described by a relaxation modulus that is the product of a time-dependent term and a straindependent term. This observation inspired White and Tokita [14] to
suggest that the memory function for a polymeric liquid might be
expressed as the product of a strain-independent function of time
and a function of strain. Wagner [15] proposed the following
formulation:
(3-63)
where h(Il' 12 ) is called the "damping function."
In this case the memory function is said to be "separable" or
factorable, and the constitutive equation becomes

Tij(t)

t
-

m(t - t')h(Il' 12 )B;j(t, t') dt '

(3-64)

00

By comparison with Equation 3-60 we see that this is a factorable


BKZ equation with the Cij term omitted. We will call this "Wagner's
equation." Since the memory function is independent of strain, it is
the same as the memory function determined in a small strain
linear viscoelastic experiment. It is thus related to the linear relaxation modulus by Equations 3-21 to 3-24.
Another way to generalize the rubberlike liquid is by using a
nonlinear strain measure, Sit, t') in place of the Finger tensor.

T;j(t)

m(t - tl)S;j(t,tl)dt '

(3-65)

-00

Seth [16] has proposed the use of a specific form of nonlinear stress
tensor. Booij and Palmen [17], however, have shown that the use of
the Seth strain measure cannot describe experimental data for a
number of melts and concentrated solutions except when the total
strain is small. They propose instead a nonlinear strain measure
that is equal to a scalar function of the Finger tensor times the
Finger tensor. However, such a model is indistinguishable from that

INTRODUCTION TO NONLINEAR VISCOELASTICITY

131

given by Equation 3-64. In other words, the damping function can


be considered to be part of a nonlinear memory function or part of
a nonlinear strain measure. In the latter case:
(3-66)

From a practical point of view this distinction is of no importance, but it is of interest to theoretical rheologists, as in the case of
a model derived from molecular dynamics, it will reflect certain
assumptions of the model. Furthermore, it may guide the modification of the model to improve its predictive abilities.
3.7.2 Determination of the Damping Function

Wagner's equation (3-64) is not a complete constitutive equation,


since it contains the unknown function, hU1 ,I2 ), which must be
determined for each polymer. However, it is useful as a tool in the
interpretation of experimental data and in the prediction of behavior in a large, rapid deformation on the basis of behavior in another
such deformation. Equations relating the damping function to several nonlinear material functions are presented in Chapter 5.
It is important to note that it is not possible to determine the
specific dependence of h on the two scalar invariants of Bij using
data from conventional experiments. This is because in simple
shear and in simple extension the two invariants cannot be varied
independently. For simple shear we showed in Section 3.4.4 that
the scalar invariants are:
(3-14)

Thus, the only independent variable is the shear strain, and all that
can be determined in a simple shear experiment is the dependence
of the damping function on 'Y.
For simple extension the first and second invariants of the Finger
tensor are given by Equations 3-15 and 3-16. In this case the
relationship between II and 12 is not so simple, but it is clear that
both invariants depend only on the Hencky strain, B. Thus, all that
can be determined in a simple extension experiment is the depen-

132

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

dence of the damping function on E. The fact that II and 12 cannot


be varied independently in either simple shear or simple extension
is illustrated in Figure 3-1.
If both h( y) and h(E) have been determined experimentally, and
empirical equations have been fitted to the data, one can usually
concoct a function h(Il,I2 ) that will encompass both of these as
special cases. However, there is nothing unique about a function
formulated in this way, and there is no guarantee that it has any
validity for flows other than simple shear and simple extension.
Furthermore, it cannot be used to predict extensional flow behavior
based only on data obtained in shear experiments or vice versa.
However, presently available evidence suggests that it is possible
to predict behavior in a variety of strain histories, as long as these
involve the same kinematics as the flow used to determine the
damping function. For example, one can often use the h( y) function determined by means of step shear strain experiments to
predict YJ +(t, y) and YJ( y). Furthermore, it is sometimes possible to
predict the first normal stress difference in any simple shear flow.
3.7.3 Separable Stress Relaxation Behavior

For a shear strain of magnitude y, if the memory function is


separable, the shear stress is given by:

a{t)

yh{y)G{t)

(3-67)

From Equation 3-67 the nonlinear relaxation modulus is:

G{t,y)

h{y)G{t)

(3-68)

where G(t) is the linear relaxation modulus. Thus, separability of


the memory function implies separability of the relaxation modulus.
Any constitutive equation that predicts a separable nonlinear relaxation modulus (Equation 3-68) will provide a prediction of the
function h( y). Khan and Larson [18] give the damping functions
predicted by a number of empirical constitutive equations.
The first normal stress difference is:
(3-69)

INTRODUCTION TO NONLINEAR VISCOELASTICITY

133

From Equations 3-67 and 3-69 we see that:

(3-70)

This implies that the Lodge-Meissner relationship, Equation 3-35,


may be valid even when the strain magnitude is beyond the range
of applicability of the rubberlike liquid model. In fact, the LodgeMeissner relation has been found to be obeyed by several molten
polymers up to moderate values of the strain.
As will be shown in Chapter 5, there are also relationships
between h(,,) and the material functions that can be determined in
several other shear flows. However, these relationships are significantly more complex than Equation 3-68 and require the
differentiation of data, and step strain experiments provide the
most straightforward method for the determination of h(,,).
A simple relationship also exists between the tensile stress relaxation function and h( e):

(3-71)
The experimental determination of (Tit), however, is quite difficult,
and few attempts have been made to date to determine h(e) from
tensile step-strain experiments [19].
Equations 3-67 and 3-69 imply that if the logarithm of the
nonlinear modulus, determined by measuring (T(t)/" or Nit)/,,2,
is plotted as a function of time (or log time), the curves for various
strains can be brought together by means of vertical shifts. Figure
3-3 is a plot of relaxation modulus data obtained by Laun [20] for
low density polyethylene "Melt I." He found that the data for (T(t)
and N 1(t)/,,2 lay on the same curves for all values of ". The curves
were found to be superposable by means of a vertical shift, and the
value of the damping function for each strain could be determined
from the shift factor.
Separability is not a universally observed phenomenon; a thorough discussion of this question is presented in Chapter 5.

134

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


105~

______________________________________,

10 4

<?
~

.....

"-.

10 3

' -'

"0

c:

as

;:::

10 2

(5

101

TIME (5)

Figure 3-3. Relaxation modulus data (open symbols) and N 1(t)/y2 (closed symbols) for
LDPE "Melt I" as determined by Laun [20] for various strain magnitudes. Starting with the
uppermost symbols (circles), and moving down, the strains are: 0.2, 0.6, 1.9, 6.0, 9.5, 15.3,
18.7, 22.4 and 30.9. The solid curves were obtained by vertical shift of the linear modulus,
G(r). Adapted from Ref. 20. Copyright 1978 by Steinkopff Verlag. Reprinted by
permission.

3.7.4 Damping Function Equations for Polymeric Liquids

When experimental step strain data are found to be separable, the


damping function is a convenient tool for the representation of
the data. We review here forms that have been proposed for the
functions h( y), h(c) and h(Ip 12 ),
3.7.4. 1 Damping Function for Shear Flows

Using Meissner's experimental results for start-up of steady simple


shear of LDPE "Melt I" at 150C [21], Wagner [15] was able to fit
the data fairly well by means of a simple exponential damping

INTRODUCTION TO NONLINEAR VISCOELASTICITY

135

function:
hey)

(3-72)

exp( -ny)

Wagner found that n = 0.143 gave the best fit of the data.
Osaki [22] was able to fit his shear stress relaxation data for a
polystyrene solution by means of a sum of two exponential functions:
hey)

a[exp( -n1y)] + (1 - a)exp( -n 2 y)

(3-73)

Laun [20] determined the damping function for LDPE "Melt I"
by use of the relaxation modulus data shown in Figure 3-3. The
solid lines were obtained by vertical shift of the linear relaxation
modulus, showing that Equation 3-68 is obeyed. His curve of h( y ),
obtained from the vertical shift factors, is shown in Figure 3-4. For
values of y up to 10, Laun was able to fit his experimental h( y )
curves by means of a single exponential function (Equation 3-72)

-;:
~

z0

i=

u 0.1
z
:J
LL

(!)

a::

:::E

0.01
0

10

20

30

SHEAR STRAIN, y

Figure 3-4. Damping function for shear, h( r), for LDPE "Melt I" determined by Laun [20]
from relaxation modulus data shown in Fig. 3-3. Also shown are best fits of Equations 3-72
(dashed line) and 3-73 (solid line). Adapted from Ref. 20. Copyright 1978 by Steinkopff
Verlag. Reprinted by permission.

136

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

with n = 0.18. We note that this is somewhat different from the


value of 0.143 determined by Wagner for the same polymer. However, G(t, y) is more sensitive to the detailed nature of the damping
function than 7J +(t, y), which Wagner used to determine h( y). As
can be seen in Figure 3-4, a single value of n cannot accommodate
Laun's data over his entire range of shear strain. However, he was
able to obtain a good fit of the entire curve by the use of Osaki's
function (Equation 3-73) with
a = 0.57

n 1 = 0.31

n z = 0.106

This function is also shown in Figure 3-4.


There is some evidence that the damping function is independent
of temperature, and if this is true it implies that the time-temperature superposition procedures that are valid for linear viscoelastic
properties will also be valid for any nonlinear property governed by
Wagner's equation.
The exponential damping functions of Wagner (Equation 3-72) or
Osaki (Equation 3-73) can be criticized on the grounds that they
indicate nonlinear behavior at all nonzero values of the shear
strain. In Chapters 4 and 5 it is shown that the prediction of some
of the material functions of simple shear are not very sensitive to
the precise form of h( y). For this type of calculation, the exponential functions are attractive, because they lead to simple, closed
forms for the equations governing these material functions. However, the absence of a finite region of applicability of the theory of
linear viscoelasticity is clearly at variance with observations for
polymeric liquids, and this aspect of the exponential damping
functions is therefore somewhat unsatisfactory.
Zapas [23] interpreted his data for a polyisobutylene solution in
terms of the BKZ theory [12], but the shear stress results can be
equally well represented by Wagner's equation, with

1
h( y) = 1 + ayZ

(3-74)

INTRODUCTION TO NONLINEAR VISCOELASTICITY

137

Zapas found that a = 2/9 gave a good fit to his polyisobutylene


data, while Adams and Bogue [24] found that a = 1/9 gave a
satisfactory fit of their viscosity data for a solution of polystyrene in
Aroelor.
Larson [25] reports that Equation 3-74 closely approximates the
damping function predicted by the molecular theory of Doi and
Edwards (see Section 3.8) if a is set equal to 0.2. This theory
applies only to linear, mono disperse polymers, but Larson suggests
that Equation 3-74 can also describe the behavior of polydisperse
and branched polymers if a is allowed to take on values between
0.04 and 0.2.
Soskey and Winter [26] have proposed a generalization of Equation 3-74 as follows:

h( 'Y)

1
1 + a'Yb

(3-75)

They found that their step strain data for both a LDPE and a broad
molecular weight polystyrene could be represented well by Equation 3-75 up to 'Y = 25 with the parameters shown in Table 3-1.
Table 3-1. Parameters of Equation 3-75 [26]
a
b

HOPE

POLYSTYRENE

0.172
1.39

0.302
1.57

Petrie [27] has presented tabulations and plots of a number of


forms for h( 'Y) that have been proposed by various researchers, and
Booij and Palmen [17,28] have assembled several additions to
Petrie's list. 7 Leblans et al. [29] have presented graphs of the
damping functions determined experimentally for several commercial polyolefins. 8 They found that the damping function was a
function of chemical structure but not strongly influenced by the
7The equations and graphs presented by Petrie [27] give the nonlinear strain, S12' as a
function of k, where k = ')' and S12 = 0.5,),h(,),), so that h(')') = 2S12 /,),. Booij and Palmen
[17] give equations for 2Sxy , which is equal to ')'h.
SIn fact, their plots show the nonlinear strain function, S12' as a function of the shear strain.
The damping function is equal to 2S12/,),.

138

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

o-------------------------------------~
-1

-2
~

'"
~
E

(5-')=1
!J
0.1
0.03 L\
0.01 0

-3

-4

-5
0

STRAIN, e

Figure 3-5. Values (of the natural logarithm) of the damping function h(e) for LDPE "Melt
I" as calculated by Wagner [30] using 150C tensile start-up flow data of Meissner (31). Curve
is Wagner's fit of Equation 3-76 with m = .30 and ex = .0025. Adapted from Ref. 30.
Copyright 1978 by Elsevier Science Publishers. Reprinted by permission.

molecular weight distribution. However, much remains to be learned


about the effect of chemical and physical structure on the damping
function.
3.7.4.2 Damping Function for Simple Extensional Flows

Turning now to h(), which can be determined in simple extension,


it has already been noted that step tensile strain experiments are
difficult to perform. Wagner [30] derived an equation (shown in
Chapter 6) that can be used to calculate h() from data obtained in
a tensile start-up experiment. Using data of Meissner [31] for LDPE
"Melt !" he obtained the h() values plotted in Figure 3-5. He was
able to fit the data reasonably well using the following function:
h()

{a[exp(2)] + (1 - a)exp(m)}-l

(3-76)

INTRODUCTION TO NONLINEAR VISCOELASTICITY

139

A plot of this function is also shown in Figure 3-5. While the


agreement is encouraging, it is to be noted that this function does
not have the expected limiting behavior at small values of e and
that the experimental data are limited to quite low Hencky strain
rates, usually below 1 s -1.
Leblans et al. [19] analyzed data from several types of simple
extensional flow for both high and low density polyethylenes. It was
possible to interpret all of the results for the LDPE in terms of a
single damping function, but for the HDPEs, the damping function
appeared to depend on the strain rate.
3.7.4.3 Universal Damping Functions

Given only h( y) and h(e), as determined in simple shear and


simple extension respectively, it is not possible to determine a
unique function hU1,12 ). For example, Wagner [15], in discussing
the single exponential Equation 3-72 for the damping function in
shear, pointed out that there are an infinite number of functions
hU1,12 ) that will reduce to Equation 3-72 for shear flow, the
simplest of which are:

(3-77)
and

(3-78)
These equations have not been found suitable for the prediction of
the damping function in simple extension.
However, it is an interesting challenge to look for the simplest
possible empirical function hU1, 12 ) that is consistent with the
observed forms of both h( y) and h(e). Wagner [11] found that he
could bring together LDPE data from both shear and extension by
plotting h as a function of the invariant, 1, where

(3-79)
Thus, this definition can be used to generalize a function determined either from h( y) data or from h(e) data. However, the
parameter f3 cannot be determined by means of shear experiments

140

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

alone, since for simple shear, II

12 , and:
(3-80)

Papanastasiou et al. [32] found that they could represent shear


and extensional flow data for polydimethylsiloxane, low density
polyethylene [20] and polystyrene [33] by a damping function that
can be written in terms of Wagner's general invariant, I, defined by
Equation 3-79:

h(I)

[1 + a(I - 3)]-1

(3-81)

The only notable inconsistency in fitting experimental data to this


equation arose in the case of the first normal stress difference
during start-up of steady simple shear. The data show a mild
overshoot, while the use of Equation 3-81 predicts no overshoot.
The values of the two parameters that were found to give a good fit
of all the data are given in Table 3-2. It can be seen that a is
somewhat more material dependent than (3.
Table 3-2. Parameters for Equation 3-81 [32]
POLYMER

polydimethylsiloxane
low density polyethylene
polystyrene

f3

0.20
0.070
0.095

0.021
0.018
0.022

For simple shear Equation 3-81 is equivalent to Equation 3-74.


Thus, the values of a shown in Table 3-2 can be compared with the
value of 0.22 reported for a polyisobutylene solution [23] and the
value of 0.11 reported for a polystyrene solution [24]. Figure 3-6 is a
plot of Equation 3-74 for the three values of a shown in Table 3-2.
We note once again that this function has the desired property that
nonlinearity appears as a second order effect.
Laun [10] has argued that separability is only approximately
obeyed in most cases, and that the use of complicated equations for
h( 'Y) is not justified. While the simple exponential form (Equation

INTRODUCTION TO NONLINEAR VISCOELASTICITY

141

0.9

;::

0.8

0.7
Z
0
i= 0.6

()

u.. 0.5

:::l
(!)

Z 0.4
ii:
::2
0.3

0.2
0.1
0

10

SHEAR STRAIN, y

Figure 3-6. Plot of Equation 3-74 for the values of a shown in Table 3-2: curve I, a
curve 2, a = .07, curve 3, a = .095.

.2;

3-72) may not be suitable for use with narrow MWD linear polymers, commercial thermoplastics are rarely if ever of this type.

3.7.5 Interpretation of the Damping Function in Terms of Entanglements

When the damping function is equal to 1, we recover the rubberlike


liquid model, which is believed to be a good representation of small
deviations from linear viscoelasticity. For deformations in which the
magnitude and rate of the deformation are fairly large, however, it
is always found that the damping function decreases as the strain is
increased. It was mentioned in Section 3.5.2 that Lodge has shown
that the rubberlike liquid model can be derived on the basis of a
network theory for entangled polymeric liquids. In Lodge's network
theory the entanglement density is assumed to be unaffected by the
deformation. This fact inspired Wagner and Stephenson [34,35] to
suggest that the damping function can be interpreted as the proba-

142

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

bility that a network strand created at time t' will survive the effects
of deformation until time t.
3.7.5.1 The Irreversibility Assumption

When separability is found to be valid for a particular polymer in a


step strain experiment, i.e., when Equation 3-68 is found to be
valid, then the damping function can be determined, and the
response of the melt to other shear histories can be calculated using
Equation 3-64, together with Equation 3-2.
This procedure is found to give fairly good results as long as the
shear strain is always in the same direction. However, when the
flow reverses direction, Equation 3-64 is found to be inadequate to
describe the behavior. Wagner and Stephenson [34,35] use the
interpretation of hU1 ,I2 ) as the network survival probability to
explain this phenomenon. They note that since h decreases as II or
12 increases, h decreases as the strain increases. If the direction of
the strain is now reversed, II and 12 will decrease, and hU1, 12) will
increase, which implies that the deformation is promoting the creation of entanglements. But this is inconsistent with the physical
picture of an entanglement network, in which entanglements (and
network strands) are lost through relaxation (thermal) processes
and through strain-induced disentanglement, while they are reformed only as a result of Brownian (thermal) motion and not
through a strain-related mechanism.
This interpretation led Wagner and Stephenson to propose an
"irreversibility assumption," which says that Equation 3-64 is only
valid for "nondecreasing deformations" and that in a "decreasing"
deformation, hU1, 12 ) should be replaced by a damping functional,
HU1, 12 ), which is the minimum value of h in the time interval
between t' and t.
(3-82)

They tested this concept by examining the results of experiments


involving recoverable strain in both simple shear and simple extension [34,36]. They found that the recoverable strain in extension

INTRODUCTION TO NONLINEAR VISCOELASTICITY

143

6r-------------------------~
5

;;: 4

~ 3

a:

w
~ 2

ow

a:

TOTAL STRAIN,

4
E =

it

Figure 3-7. Recoil during tensile start-up flow for LDPE Melt I at 6 = 1 S-1 and 150C;
comparison of predictions of Wagner's equation with (curve 2) and without (curve 1) the
irreversibility assumption. Data are those of Meissner [31]. Adapted from Ref. 34. Copyright
1979 by The Society of Rheology. Reprinted by permission of John Wiley & Sons, Inc.

could not be described in terms of a damping function but that the


use of the irreversibility assumption provided a much better fit of
the data. This is shown in Figure 3-7. In the case of constrained
recoil during start-up of steady simple shear, however, the use of
the damping function gave a prediction only moderately larger than
the observed recoverable shear, and the use of the damping functional gave only a slightly better fit of the data. This is shown in
Figure 3-8.
A shear test that is more sensitive to the use of HUt, 12 ) in place
of hUt, 12 ) is a double step strain in which the second step is in the
direction opposite to that of the first. In a double-step reversing
strain test, the predictions of Wagner's equation with and without
the reversibility assumption are quite different. Figure 3-9 shows
data for a LLDPE as well as the predictions of both equations [37].
The first strain has a value of 5, while the second occurs 0.5 seconds
later and has a value of - 1. The time in Figure 3-9 is measured
from the application of the second strain step. The damping function used for the prediction was that given by Equation 3-74 with

144

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

~ 10'
a:

tn

a:

en
~ 10
III

a:

o>
()

a:

10-'~--------~--------~------~

10-'

10

10'

10 2

SHEAR STRAIN, y

Figure 3-8. Recoil during start-up shear flow of a LDPE melt at y = 2 s-\ comparison of
predictions of Wagner's equation with (curve 3) and without (curve 2) the irreversibility
assumption. Curve 1 is the prediction of the rubberlike liquid model. Data are those of
Meissner. Adapted from Ref. 34. Copyright 1979 by The Society of Rheology. Reprinted
by permission of John Wiley & Sons, Inc.

a = 1/2. Clearly the irreversibility assumption leads to a prediction


that is much more in agreement with the data than the basic form
of Wagner's equation.
It is important to note that even in the case of deformations in
which the irreversibility assumption must be invoked, as long as
separability is valid, the damping function, h(Ip 12 ) still contains a
complete description of the nonlinearity of the viscoelastic behavior
of the material.
3.7.6 Comments on the Use of the Damping Function

There exists at present no useful general theory of nonlinear


viscoelastic behavior. This means that our understanding of such
behavior is entirely empirical, i.e., based on a collection of
individual observations. Given this situation, any techniques that
help us to present and compare such observations in meaningful
ways are of value. The use of the damping function is one of the
few techniques available. First, it is useful for bringing together
data for various values of strain or strain rate and presenting these
in terms of a single curve of h( 'Y) or h(E). Secondly, even when

INTRODUCTION TO NONLINEAR VISCOELASTICITY

145

8
6

<U

Q.

6
0

iii
(J)
w

a:

I-

..

(J)

a:

<

(J)

- 60.1
----------~---*----------~--~--------~
0.5
10
5

50

T IME (5)

Figure 3-9. Stress relaxation following a double step, reversing shear strain with 'Yl

5 (at

t = 0) and 'Y2 = -1 (at t = .5 s). Comparison of experimental data with the predictions of

Wagner's equation with (curve 1) and without (curve 2) the irreversibility assumption.
Adapted from Ref. 37. Copyright 1986 by The Society of Rheology. Reprinted by
permission of John Wiley & Sons, Inc.

separability fails so that the data cannot be brought onto a single


master curve, we at least have a procedure for categorizing nonlinear behavior, i.e., we can report that the nonlinear behavior is
either separable or nonseparable, for a given material over a given
range of times and strain magnitudes. This suggests the following
classification structure:
1. Linear viscoelasticity

(obeys Boltzmann superposition principle).


2. Finite linear viscoelasticity
(obeys the rubberlike liquid equation).
3. Separable nonlinear viscoelasticity.
4. Nonseparable nonlinear viscoelasticity
(the most general case).

146

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

At the same time, it must be remembered that Wagner's equation is not a constitutive equation, and it cannot be used to predict
the response of a melt to deformations with kinematics different
from that of the deformation used to determine the damping
function. Furthermore, the damping function is material specific,
and this means it is not possible to predict the behavior of one
polymer on the basis of experiments on another.

3.8 MOLECULAR MODELS FOR NONLINEAR VISCOELASTICITY

If the modified Rouse model of Chapter 2 is extended to predict

the response of a melt to large strains, some nonlinear effects make


their appearance. For example, the first normal stress difference is
nonzero in steady shear and in step shear. However, there is no hint
of the most striking nonlinear phenomena that have been observed
experimentally. These are clearly manifestations of entanglement
coupling, and to gain any understanding of these at the molecular
level, a model that incorporates entanglements in some manner
must be used.
The concept of reptation was introduced in Section 2.11.2 to
explain the remarkable effects of entanglement coupling on the
linear viscoelastic behavior of high molecular weight polymers.
Although the calculations are more difficult when one wishes to
extend this concept to the prediction of nonlinear behavior, some
progress has been made in this area, and this will be reviewed here.
In Chapter 2, it was noted that according to the reptation theory
of Doi and Edwards [3], two characteristic times govern linear
viscoelastic behavior. First is the equilibration time, Ae , which is the
time scale for the reorganization of chain segments over lengths less
than a, i.e., between entanglement points. Relaxations occurring
during this very short time scale can be described by the Rouse
model, as they are not impeded by the "tube" or entanglement
network. Over longer time scales, however, the tube severely restricts the motions of a molecule, and the Rouse model is no longer
valid. The Doi-Edwards theory can be used to treat that portion of
the relaxation process that occurs over time scales longer than Ae'
To accomplish this, an additional time scale was introduced and

INTRODUCTION TO NONLINEAR VISCOELASTICITY

147

(5'

Figure 3-10. Nonlinear stress relaxation modulus predicted by the theory of Doi and
Edwards for Ad/AR = 100. At large strains, the retraction mechanism for relaxation becomes
important at intermediate times, but disengagement is still the final relaxation process.
Adapted from M. Doi, I. Polym. Sci. 18:1005 (1980). Copyright John Wiley & Sons, Inc.

this was Ad' the disengagement time, which governs the motion of a
molecule out of its tube.
To understand the response of a melt to large, rapid deformations, a third relaxation mechanism must be taken into account, and
this is contour length relaxation. This is the retraction of a molecule
within its tube back to its equilibrium length, following a deformation of the melt and thus of the tube. For a linear, monodisperse
polymer, the time scale for this process is the longest Rouse
relaxation time, AR'
The Doi-Edwards theory [3] can be used to predict the straindependent relaxation modulus G( ,,/, t) for t > Ae , and the results
are shown in Figure 3-10. For small values of ,,/, the linear viscoelastic prediction is approached, and the modulus becomes
independent of strain. As the strain is increased, however, the
relaxation modulus begins to be strongly affected by the retraction
process, and we see a fairly distinct separation between the relaxation due to retraction and that due to disengagement, which
occurs at longer times. The general behavior of the curves is in

148

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

agreement with the predictions (see Section 5.2.2).


For times longer than the retraction time the theory predicts that
the modulus is separable, i.e., that:

G(y,t)

h(y)G(t)

(3-68)

The theory also predicts that the first normal stress difference
during relaxation is given by:

(3-70)
This is the Lodge-Meissner relationship, which is also predicted by
the rubberlike liquid model and is in agreement with many experimental observations.
The prediction of the tensile relaxation modulus has been worked
out in closed form [3]. Unlike the shear modulus, it is not separable
into strain and time-dependent factors.
The theory has been used to predict responses to a few other
flows, and this will be discussed in Chapter 5. However, the calculations are difficult, and except for a few simple cases, these
predictions make use of the "independent alignment" assumption,
described in the next section.
3.8.1 The Doi-Edwards Constitutive Equation

In order to derive a constitutive equation convenient for the calculation of responses to a wide variety of deformations, Doi and
Edwards found it necessary to make two simplifying assumptions.
One of these is that the molecule does not stretch within its tube
when the melt is subjected to deformation. In other words, the
retraction (contour length relaxation) process occurs instantaneously. This assumption is valid when the time scale of the
experiment (or the reciprocal of the strain rate) is greater than AR"
The second assumption is also made necessary by complications
arising from the retraction of a molecule within its tube. In particular, because of retraction, it is not possible to associate a particular
strand (a group of segments extending between two entanglements)
with a particular portion of the tube. To avoid this complication,
Doi and Edwards made what they called the "independent align-

INTRODUCTION TO NONLINEAR VISCOELASTICITY

149

ment" (IA) assumption. The effect of this assumption is to limit the


theory to the prediction of mildly nonlinear effects. Recoil and the
response to reversing double-step strains are not correctly
predicted.
With these two simplifying approximations, Doi and Edwards
were able to derive a constitutive equation. The result is a separable BKZ equation, whose general form is given by Equation 3-60.
The potential function, U(/I' [2)' is not given in an explicit form by
the theory, although approximate analytic expressions have been
proposed [2, 38, 39].
For steady simple shear, the Doi-Edwards constitutive equation
predicts that at high shear rates (yA d 1) the viscosity is given by:
. \

T'J = T'Jo ( ylld

) -3/2

(3-83)

This implies that the shear stress decreases as the shear rate
increases. Since the stress initially increases with y, this prediction
implies an unstable situation in which there is more than one
possible shear rate corresponding to a given shear stress. While it
has been suggested that this provides an explanation of the irregular flow that sometimes occurs in capillary flow [40] no polymeric
liquid has been clearly shown to have the same value of the shear
stress for two values of the shear rate in steady simple shear. The
theory also predicts that the first normal stress coefficient decreases
sharply with y at high shear rates ('1'1 a y-2). As in the case of the
viscosity, experiments indicate a milder decrease. One factor contributing to these incorrect predictions is the very narrow relaxation
spectrum predicted by the original Doi-Edwards theory.
The predictions of the Doi-Edwards equation in shear, except for
the second normal stress difference, are equivalent to those of
Wagner's equation with specific forms for the memory function
and the damping function. The theory does not provide an analytic
form for h( y), but Larson [2] has noted that Equation 3-74 is
a close approximation if a = 0.2. The theory predicts that h( y) is
independent of chemical structure. Like Wagner's equation, the
Doi-Edwards constitutive equation cannot describe responses to
deformations involving flow reversal. Thus, recoil phenomena and
responses to reversing step strains are incorrectly predicted.

150

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

3.9 STRONG FLOWS: THE TENDENCY TO STRETCH AND ALIGN MOLECULES

We have seen that outside the regime of linear viscoelastic behavior, each type of flow reveals a different aspect of the nonlinear
rheological behavior of a polymeric liquid. It would be very useful
to know something about the molecular mechanisms underlying this
behavior. For example, it would be of interest to know to what
extent a particular type of flow tends to produce disentanglement,
the stretching of molecules, and the alignment of molecules with
each other. Ultimately, this might permit us to predict the general
features of rheological behavior on the basis of the molecular
structure of a polymer.
The characteristics of a flow that are of importance with regard
to the type of molecular effects produced are the strain magnitude
history and the kinematics. By strain magnitude history is meant the
way that the strain varies with time. For example, start-up of simple
shear brings out different features of the behavior than large
amplitude oscillatory shear. By kinematics is meant the geometric
nature of the deformation, which is indicated by the form of the
strain tensor. For example, simple shear is kinematically quite
different from simple extension. It is not at all clear at this time
exactly how these aspects of a flow influence changes in molecular
physical structure in a melt, but some ideas have been advanced.
Tanner and Huilgol [1,41] coined the phrase "strong flow" to
describe a deformation that has a strong tendency to stretch out
molecules. They showed that steady simple extension is strong in
this sense, while steady simple shear is "weak." Doshi and Dealy
[42] expanded on this idea and noted that not all shear flows are
weak. In particular they showed that exponential shear (see
Section 5.9) is a strong flow.
Larson [43] has suggested that a strong flow does not necessarily
generate a high degree of molecular alignment. He proposes, as a
measure of the tendency of a flow to generate alignment, the
difference between the two principal invariants of the Finger strain
tensor (see Section 3.4.4). Specifically, he proposes the following
classification:
I} - 12 > 0

High tendency to align molecules

I} - 12

Neutral tendency to align molecules

I} - 12 < 0

Little tendency to align molecules

INTRODUCTION TO NONLINEAR VISCOELASTICITY

151

We note that this is a purely kinematic criterion that does not take
into account the strain magnitude history.
Samurkas et al. [44] noted that both simple shear and planar
extension have the property that II = 12 and thus according to
Larson's scheme should have the same aligning tendency. Furthermore, they noted that by using exponential shear, rather than a
constant shear rate, simple shear can be a strong flow. Thus, they
compared the response of the same melt to exponential shear and
planar extension. However, the damping functions necessary to fit
the two results were quite different, and this throws considerable
doubt on the usefulness of these schemes for classifying flows.
To complicate the picture further, Meissner and Demarmels
[45,46] have generated well-controlled laboratory flows in which the
kinematics change with time so that there is the additional feature
of the kinematic history. Much remains to be learned about how all
these aspects of a deformation influence the physical structure of a
polymeric liquid.
REFERENCES
1. R. I. Tanner, Engineering Rheology, Oxford University Press, Oxford, 1985.
2. R. Larson, Constitutive Equations for Polymer Melts and Solutions, Butterworths, Boston, 1988.
3. M. Doi and S. F. Edwards, The Theory of Polymer Dynamics, Oxford University
Press, Oxford, 1986.
4. R. B. Bird, O. Hassager, R. C. Armstrong and C. F. Curtis, Dynamics of
Polymeric Liquids, Volume 2, Kinetic Theory, Second Edition, John Wiley &
Sons, NY, 1987.
5. K. N. Sawyers, J. Elasticity 7:99 (1977).
6. A. S. Lodge, Elastic Liquids, Academic Press, NY (1964).
7. A. S. Lodge, Trans. Faraday Soc. 52:120 (1956).
8. A. S. Lodge and J. Meissner, Rheol. Acta 11:351 (1972).
9. A. S. Lodge, Journal Non-Newt. Fl. Mech. 14:67 (1984).
10. H. M. Laun, J. Rheol. 30:459 (1986).
11. M. H. Wagner, Rheol. Acta 18:33 (1979).
12. B. Bernstein, E. A. Kearsley and L. J. Zapas, J. Res. Nat. Bur. Stds. 68B:103
(1964).
13. R. I. Tanner, J. Rheol. 32:673 (1988).
14. J. L. White and N. Tokita, J. Phys. Soc. Japan 22:719 (1967).
15. M. H. Wagner, Rheol. Acta 15:136 (1976).
16. B. J. Seth in: M. Reiner and D. Abir eds., Second Order Effects in Elasticity,
Plasticity and Fluid Dynamics, p. 162, Macmillan, NY, 1964.
17. H. C. Booij and J. H. M. Palmen, Rheol. Acta 21:376 (1982).

152

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

18. S. A. Khan and R. G. Larson, J. Rheol. 31:207 (1987).


19. P. J. R. Leblans, J. Sampers and H. C. Booij, J. Non-Newt. Fl. Mech. 19:185
(1985).
20. H. M. Laun, Rheol. Acta 17:1 (1978).
21. J. Meissner, Rheol. Acta 10:230 (1971).
22. K. Osaki, Proc. VIIth Intern. Congr. Rheol., p. 104, Gothenburg, 1976.
23. L. J. Zapas, J. Res. Nat. Bur. Stds. 70A:525 (1966).
24. E. B. Adams and D. C. Bogue, A.I.Ch.E.J. 16:53 (1970).
25. R. G. Larson, J. Rheol. 29:823 (1985).
26. P R. Soskey and H. H. Winter, J. Rheol. 28:625 (1984).
27. C. J. S. Petrie, J. Now-Newt. Fl. Mech. 5:147 (1979).
28. H. c. Booij and J. H. M. Palmen, J. Non-Newt. Fl. Mech. 23:189 (1987).
29. P. J. R. Leblans, J. Sampers and H. C. Booij, Rheol. Acta 24:152 (1985).
30. M. H. Wagner, J. Non-Newt. Fl. Mech. 4:39 (1978).
31. J. Meissner, Rheol Acta 10:230 (1971).
32. A. C. Papanastasiou, L. E. Scriven and C. W. Macosko, J. Rheol. 27:387
(1983).
33. H. M. Laun, M. H. Wagner and H. Janeschitz-Kriegl, Rheol. Acta 18:615
(1979).
34. M. H. Wagner and S. E. Stephenson, J. Rheol. 23:489 (1979).
35. M. H. Wagner and S. E. Stephenson, Rheol. Acta 25:463 (1979).
36. M. H. Wagner and J. Meissner, Makrornol. Chemie 18:1533 (1980).
37. R. G. Larson and V. A. Valesano, J. Rheol. 30:1093 (1986).
38. P. K. Currie in Rheology (Proc. VIn Intern. Congr. Rheol., Naples) Ed. by G.
Astarita et aI., Vol. 1, p. 357, Plenum Press, New York, (1980).
39. R. G. Larson, in Rheology (Proc. IX Intern. Congr. Rheol., Acapulco), Ed. by
B. Mena et aI., Vol. 1, p. 467, UNAM, Mexico City (1984).
40. Y. H. Lin, J. Rheol. 29:605 (1985).
41. R. I. Tanner and R. R. Huilgol, Rheol. Acta 14:959 (1975).
42. S. R. Doshi and J. M. Dealy, J. Rheol. 31:563 (1987).
43. R. G. Larson, J. Non-Newt. Fl. Mech. 23:249 (1987).
44. T. Samurkas, R. G. Larson and J. M. Dealy, J. Rheol. 33:559 (1989).
45. J. Meissner, Ann. Rev. Fluid Mech. 17:45 (1985).
46. A. Demarmels and J. Meissner, Coli. Polym. Sci. 264:829 (1986).

Chapter 4
Steady Simple Shear Flow
and the Viscometric
Functions
4.1 INTRODUCTION

Steady simple shear is of central importance in applied rheology for


two reasons. First, it is the flow that is by far the easiest to generate
in the laboratory. Therefore, the data most often reported are
based on this flow. Secondly, a number of processes of industrial
importance, particularly extrusion and flow in many types of die,
approximate steady simple shear flow. For these reasons it seems
appropriate to devote an entire chapter to this subject.
4.2 STEADY SIMPLE SHEAR FLOW

Simple shear flow was introduced in Section 1.4, and we will review
its essential features here. Referring to Figure 4-1, we see that this
flow is generated by the rectilinear motion of one flat plate relative
to another, where the two plates are parallel and the gap between
them, h, is constant with time. This flow is completely described by
giving the shear strain, y, as a function of time, where y is defined
by Equation 1-10.
y

aXjh

(1-10)

The shear rate is the derivative of this quantity with respect to time
and is thus the velocity of the moving plate divided by the plate
153

154

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

~f----~~~~~;=~~V

....L~~;:;;:~~

Figure 4-1. Simple shear flow. The upper plate moves at velocity V in the
lower plate is fixed.

XI

direction. The

spacing:

Y == IY2l1 = Iyn! = IVI/h

(4-1)

Lodge [1, p. 62] shows that for simple shear flow, the only nonzero
components of the stress tensor are those shown in Equation 1-29:

(1-29)

But the stress tensor is symmetric, and if we use the symbol u to


mean the magnitude of the shear stress:
(4-2)

As explained in Section 1.8, the magnitude of the normal stress


components have no rheological significance for an incompressible
fluid, because if all three components are changed by the same
amount, no deformation will be generated. However, differences
between these components can generate deformation and are rheologically significant. It is customary to define the first and second
normal stress differences as follows:

(4-3)
(4-4)
Thus, for simple shear flow, there are only three rheologically
significant features of the stress tensor: u, N l , and N 2

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

155

Steady simple shear is a simple shear flow that has been carried
out at a constant shear rate for a sufficient length of time that the
stresses have reached steady values. In this case, the stresses are
functions only of the shear rate, y: a( y); N l ( y); N 2 ( y). For a
Newtonian fluid, the shear stress is proportional to shear rate,
a = TJY, and Nl = N2 = O. The shear stress is not, in general,
proportional to the shear rate, and the first and second normal
stress differences are not equal to zero. At very small shear rates,
however, it is usually found that the shear stress does become
proportional to shear rate, while the first normal stress difference
becomes proportional to the square of the shear rate. This is the
type of behavior predicted by the Lodge rubberlike liquid model
presented in Chapter 3. This low shear rate behavior has inspired
the definition of the following material functions for steady simple
shear:
The Viscosity

TJ == alY

(4-5)

The First Normal Stress Coefficient

(4-6)
The Second Normal Stress Coefficient

(4-7)
These three functions are, in general, functions of the shear rate
and are the three measurable material functions of steady simple
shear flow.
4.3 VISCOMETRIC FLOW

Steady simple shear is a uniform deformation, i.e., each fluid


element undergoes exactly the same deformation, and the extra
stresses are independent of position in space. There are also
nonuniform flows for which the three material functions defined
above govern the behavior of the fluid. Such deformations are
called "viscometric flows." We will define a viscometric flow as one
which, from the point of view of a fluid element, is indistinguishable

156

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

from steady simple shear. A more comprehensive mathematical


definition can be found in the book by Bird et al. [2, p. 157]. Thus,
while different fluid elements in the field of flow may be subject to
different shear rates, the shear rate experienced by any particular
fluid element is constant with time.
For a viscometric flow, knowledge of the three material functions
of steady simple shear suffice to predict the detailed nature of the
flow, i.e., the distributions of the velocity and extra stresses. For this
reason, these three functions are called the "viscometric functions."
Steady simple shear is the simplest possible viscometric flow. Viscometric flows are of special interest, because they can be used to
determine one or more of the viscometric functions. In addition,
some flows of practical importance are viscometric flows or close
approximations thereto.
Some examples of viscometric flows are listed below.
1. Steady Tube Flow (Poiseuille Flow)
This flow is the one most often used to determine the viscosity
of molten plastics. It also occurs whenever a melt is transported by means of a pipe or other circular channel. The
magnitudes of the shear stress and shear rate vary from zero
on the axis to maximum values at the wall.
2. Steady Slit Flow
Sometimes called plane Poiseuille flow, this flow can also be
used to measure viscosity. The shear stress and shear rate vary
from zero on the plane of symmetry to maximum values at the
walls.
3. Annular Pressure Flow
This flow occurs in the axial direction in the space between
two concentric cylinders as the result of a pressure gradient. If
the ratio of the two diameters is close to one, the velocity
distribution becomes very similar to that for slit flow.
4. Steady Concentric Cylinder (Couette) Flow
This is a "drag flow" generated by the rotation of either the
inner or outer cylinder of a concentric cylinder apparatus. It is
widely used for the measurement of viscosity in Newtonian
fluids. If the ratio of the two diameters is close to one, the
shear rate becomes nearly uniform in the annular gap containing the fluid and thus approximates steady simple shear flow.

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

157

5. Steady Parallel Disk Flow


This is a "torsional" flow generated when the fluid contained
between two parallel, coaxial disks is sheared by the rotation
of one of the disks.
6. Steady Cone and Plate Flow
This flow is only approximately viscometric, but it is of special
interest, because if the cone angle is very small, the shear rate
and shear stress are approximately uniform. All three viscometric functions can, in principle, be measured.
7. Steady Sliding Cylinder Flow
This is a drag flow generated when one of two concentric
cylinders is displaced along its own axis. If the ratio of the two
diameters is close to one, the shear rate is nearly uniform.
8. Steady Helical Flow
This is a combination of flows 4 and 7 above. The fluid is
contained in the annular space between two concentric cylinders; one of these is rotated at a constant speed while either
the same or the other cylinder is displaced along its own axis
at constant linear speed.
9. Combined Drag and Pressure Flows
If we combine steady simple shear with pressure flow in a slit,
there are two forces driving the fluid motion. Drag flow results
from the motion of one wall, while pressure flow results from
the pressure gradient. If the direction of the pressure gradient
is the same as the direction of motion of the moving wall, the
velocity profile is of one of the types shown in Figure 4-2 [3]. If
the direction of the drag flow is at an angle to the direction of
the pressure gradient, the resulting deformation is similar to
Case I

Case II

Figure 4-2. Velocity profiles for combined, uniaxial drag and pressure flow [3]. Cases I and
II: dP /dx < 0; cases III and IV: dP /dx > O. Adapted from Ref. 2. Copyright
1987 by John Wiley & Sons, Inc. Reprinted by permission.

158

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

that which occurs in the channel of a single screw extruder


[4, 5]. A similar combined flow is Couette flow with an axial
pressure gradient [6]. Here the drag flow direction is perpendicular to the pressure flow direction. When the gap is very
thin, this flow becomes equivalent to plane Couette flow with a
perpendicular pressure flow [7].
4.4 WALL SLIP AND EDGE EFFECTS

All of the analyses presented or cited above are based on the


"no-slip" assumption, that the fluid adheres to any wall with which
it is in contact and that if this wall moves, the fluid in contact with it
moves at the same speed. However, this assumption is not always
valid for molten plastics. For certain combinations of shear stress
and shear strain, the melt undergoes some type of fracture at or
near the wall and subsequently undergoes slip flow. In the case of
polyethylene this occurs at shear stresses in the neighborhood of 0.1
MPa, unless the experiment is terminated while the total shear
strain is still quite small.
Once slip flow begins, the velocity of the sliding polymer surface
relative to the wall is not known a priori. This seriously complicates
the interpretation of data to determine the viscometric functions.
Wall slip is discussed in further detail in Section 8.2.6.
Another assumption that is made in the classical analyses of
viscometric flows is that the deformation is homogeneous. In reality, an experimental apparatus is always finite in size, and the
sample has one or more exposed free surfaces. Examples are the
sample edge in a cone and plate or parallel disk rheometer. In
addition, there may be zones in the field of flow within which the
deformation differs significantly from the assumed viscometric flow.
Examples are the zones below the inner cylinder of a Couette
viscometer and at the entrance of a capillary rheometer. These end
and edge effects are sources of error in viscometric measurements,
and they will be considered in some detail in Chapters 7 and 8.
4.5 THE VISCOSITY OF MOLTEN POLYMERS

The viscosity is the rheological property that is most easily measured, and for this reason it is the one most frequently used to

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

159

characterize thermoplastic resins. Saini and Shenoy have published


compilations of viscosity-shear rate curves for polyethylene [8],
polystyrene [9], engineering thermoplastics [10] and specialty polymers [11].
Like other rheological properties the viscosity depends on the
following factors:
1. Flow conditions

a. Shear rate
b. Temperature
c. Pressure
2. Resin composition
a. Chemical structure of polymer
b. Molecular weight distribution of polymer
c. Presence of long chain branches
d. Nature and concentration of additives, fillers, etc.
Relationships between melt viscosity and all these factors were
covered in Kumar's 1980 review [12]. In the present chapter, the
emphasis is on the mathematical expressions that describe the
effects of flow conditions. The dependence on resin composition
factors is considered in greater detail in Chapter 10.
To obtain the viscosity over several decades of shear rate is a
time consuming task, usually requiring the use of both a capillary
rheometer (for high shear rate data) and a cone-plate rheometer
(for low shear rate data). For this reason, it is currently impractical
to use the complete viscosity function as a routine tool for quality
control. In its place it is customary to use either a single-point
measurement of a viscosity-related quantity, usually the melt index,
or the ratio of two viscosity-related quantities, such as the ratio of
the melt flow indexes measured at two nominal shear stresses
values. This matter is explored in detail in Chapter 8.
4.5.1 Dependence of Viscosity on Shear Rate

The viscosity of molten thermoplastics decreases sharply as the


shear rate is increased. Typical behavior is shown in Figure 4-3. At
sufficiently low shear rates, the viscosity normally becomes independent of shear rate, as shown in the inset of Figure 4-3. The constant

160

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

o~------------------------------------------~
o
Figure 4-3. Typical viscosity versus shear rate curve for a shear thinning material. On this
linear (non-logarithmic) scale, the approach to Newtonian behavior as y ..... 0 can only be
seen when the y scale is greatly expanded.

viscosity that prevails at very low shear rates is called the "zero
shear viscosity" and is given the symbol TJo. The zero shear viscosity
is an important scaling parameter, as is shown below, but for many
commercial resins, particularly those with very broad molecular
weight distributions or a high degree of long chain branching, it is
impossible to measure using presently available commercial
rheometers. This is because the shear rate at which TJ( y) levels out
is too low to be generated in these instruments.
In order to show clearly the approach of the viscosity to its
limiting, low shear rate value while also showing the high shear rate
behavior, it is customary to display the viscosity versus shear rate
behavior in the form of a plot of log( TJ) versus log( y), as shown in
Figure 4-4, which shows the data of Meissner [13] for low density
polyethylene "Melt I." On such a plot the points for high shear
rates often fall very close to a straight line. This suggests the use of
an empirical "power law" to describe the dependence of viscosity
on shear rate in this region:
TJ =Kyn-l

(4-8)

The shear stress is then given by:

(4-9)

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

10

10

-4

10

10

10'

161

10

y (s

')

Figure 4-4. Viscosity versus shear rate for several temperatures-Meissner's data for low
density polyethylene [13]. Temperatures, from top to bottom, are (oC): 115,130,150,170,190,
210, 240. Adapted from Ref. 2. Copyright 1987 by John Wiley & Sons, Inc. Reprinted by
permission.

Obviously, a Newtonian fluid is a special case for which n = 1 and


K becomes equal to the viscosity.
There are several awkward features of the power law expressed
by Equations 4-8 and 4-9.
1. The units of K depend on the value of n.
2. If the shear rate is negative, the equation does not yield a
value for the viscosity (unless n is an integer).
3. The zero shear viscosity does not appear as a parameter.
4. The equation is only valid at high shear rates.

The first three features can be eliminated by use of the following


form:

(4-10)
where A is a material constant with units of time, i.e., a characteristic time of the material. Specifically, it is the reciprocal of the shear
rate at which the calculated value of '17 becomes equal to '170'

162

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

For many polymers, Equation 4-8 (or 4-10) holds reasonably well
over much of the shear rate range of interest for processing. It has
been widely used in analyses of melt processes because of its
mathematical simplicity, and this makes it possible to derive explicit
equations describing many flow situations. However, these equations are often not useful for quantitative prediction because of the
other approximations required to obtain an explicit solution. Because of the wide availability of powerful computational facilities,
the power law no longer offers an important advantage, and more
realistic equations are now being used in many process simulations.
A number of generalized power law equations have been proposed that predict an approach to a constant viscosity at low shear
rates, thus overcoming the fourth undesirable feature of Equation
4-8. Cross [14] proposed the relationship shown in Equation 4-11.

(4-11)
We note that at low shear rates the viscosity approaches Tlo, while
at high shear rates (IAyl 1) power law behavior is predicted, with
m related to the power law index, n, as follows:

1- n

(4-12)

A closely related equation is the one proposed by Bueche and


Harding [15] to describe their data for concentrated solutions of
linear polymers (polystyrene and PMMA):

(4-13)
The time constant, A, in this equation is related to the molecular
weight [16].
Another generalized power law is that proposed by Carre au [17]:
( 4-14)

For additional flexibility in fitting the shape of the curve, Yasuda

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

163

et al. [18] added another constant:


(4-15)

This equation and its various special cases have been reviewed by
Hieber and Chiang [19], and a number of additional expressions for
the viscosity function have been compared by Elbirli and Shaw [20].
We note that all of these dimensionally consistent viscosity equations have the form of 7]( y) = 7] oF(,\,y), with F approaching unity
when the shear rate becomes small. As the shear rate increases, F
becomes much less than unity, decreasing very rapidly when y
becomes greater than 1/,\,. When y is appreciably greater than this
quantity, all of the equations approach simple power law behavior.
The main differences between them is the detailed shape of the
curvature of the viscosity curve in the neighborhood of y = 1/,\"
and for many applications these differences are not important.
Software supplied with data acquisition systems sold with capillary rheometers sometimes incorporate a logarithmic polynomial of
the form:

This is not to be recommended, as the equation is not constrained


to follow the correct general form for 7]( y) and can in fact lead to
curves having features that are quite unrealistic. An equation such
as (4-14) is a much better choice.
For linear polymers with narrow molecular weight distribution,
the viscosity curve (on a log-log plot) has a distinct region of
essentially constant viscosity as well as a well-defined power law
region, and the transition between the two occurs over about one
decade of shear rate. This is illustrated in Figure 4-5, where data
for several, narrow molecular weight distribution polystyrenes are
shown [21]. For polydisperse polymers, however, as well as for
polymers with significant long-chain branching, the transition occurs
over a much broader range of shear rates, as shown in Figure 4-4,
and the region of constant viscosity often is attained only at shear
rates below those accessible using standard commercial rheometers.

164

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 4-5. Viscosity versus shear rate for several narrow MWD polystyrenes. Molecular
weights are, from bottom to top: 4.85 X 10 4, 11.7 X 104, 17.9 X 10 4 , 21.7 X 10 4 and 24.2 X
10 4 . Adapted from Ref. 21. Copyright 1966 by Academic Press. Reprinted by permission.

The strong dependence of polymer viscosity on shear rate is


attributed to the strong effect of shearing on entanglements. The
primary evidence for entanglements is the effect of molecular
weight on "10 for linear, monodisperse polymers. Curves of loge "10)
versus 10g(M) for several polymers are shown in Figure 4-6 [22], At
low molecular weights the viscosity is proportional to molecular
weight and varies little with shear rate over a wide range of shear
rates. As the molecular weight increases, a point is reached where
"10 starts to increase more rapidly, and over a fairly narrow range of
M the slope of the loge "10) versus 10g(M) curve reaches a value of
about 3.4 or 3.5 for many polymers. At the same time it is found
that the viscosity begins to be more and more dependent on shear
rate. It is customary to associate this change in behavior with a
particular value of the molecular weight, called Me' and to call this
the "critical molecular weight for entanglement". Values of Me for
a number of polymers are given in Chapter 10.
The linear relationship between loge "10) and 10g(M) for highly
entangled polymers implies that
"10 = KM 3 .4

(4-16)

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

165

!=="
Ol

.2
+

IZ

<{

I-

en
Z

()

Polyethylene
Polybutadiene

4
CONSTANT

+ log M

Figure 4-6. Log 110 versus log M for several polymers. The data are shifted to avoid overlap.
The lines shown have slopes of 1.0 Cleft portion) and 3.4 (right portion). Adapted from Ref.
22. Copyright 1968 by Springer Verlag. Reprinted by permission.

166

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

For polydisperse materials, it is found that this continues to be


approximately valid if M is replaced by the weight average molecular weight.
( 4-17)
This relationship leads directly to a blending law for viscosity. For
example, in a binary blend of 2 monodisperse samples of the same
polymer having molecular weights MI and M 2 , the weight average
molecular weight of the blend is given by:
(4-18)
Where WI and w2 are the weight fractions of the two components
of the blend. Using Equations 4-16 and 4-17 to eliminate the
molecular weights, we have:
'Y)1/3.4
',Ob

W 'Y)1/3.4
1"01

+ W 2 'Y)1/3.4
"02

( 4-19)

This equation has been tested for blends of monodisperse [23,24]


and polydisperse [25] materials.
Vinogradov and Malkin [26, p. 178] suggest that the relaxation
times of a high molecular weight polymer is proportional to its zero
shear viscosity, and they have used this idea to show that one can
use the power law to estimate the effect of molecular weight on the
viscosity function. They begin by writing the ratio of the viscosities
of two melts, having molecular weights Ml and M 2 , at the same
shear rate, y, using the form of the power law given by (4-10):

7J(y,MI )
7J(y,M2)

r7J0(M )A(M r-

7J0(MI )[ A(Mlh 1n-l

7J0(MI )A(M1

7J0(M2)[ A(M2h 1n-l

(4-20)

Now they assume:


(4-21)
A a 710 a M

3 .5

( 4-22)

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

167

to obtain:

7]( y, M 1 )
7](y,M2 )

(4-23)

For example, if n = 0.3, then the high shear rate viscosity becomes
nearly proportional to M. It is important to recall here that this
analysis is valid only for high-molecular weight, linear, monodisperse polymers in the power law viscosity region.
In order to explain the decrease in viscosity with shear rate in
terms of the classical theory of viscoelasticity, this decrease has
been associated with a "truncation" of the relaxation spectrum.
From the Boltzmann superposition principle, Equation 2-18 indicates that:
7]0 =

1o G( s) ds
00

(4-24)

This can be rewritten in terms of the spectrum function, H(A) by


use of Equation 2-29, and the result is:
(4-25)
This equation is based on the theory of linear viscoelasticity, and it
thus provides no information regarding the dependence of viscosity
on shear rate. However, it has inspired the notion that during a
viscometric flow a polymer has a steady state distribution of molecular conformations and a constant entanglement density, and thus a
well-defined but shear rate-dependent relaxation spectrum. We
then modify Equation 4-25 as follows:
(4-26)
Now the reduction in entanglement density at high shear rates is
associated with the reduction of that portion of the spectrum
function corresponding to the longest relaxation times. A simple
model of this process is that as y increases, the H(A) function is
truncated at ever decreasing values of A, thus decreasing the value

168

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

of the integral and of T/. Experimental evidence indicates that this


is not quantitatively correct, but it has proven to be a useful
concept. Tanner and Simmons [27] have also discussed the effect of
shear rate on the relaxation spectrum function.
It is important to keep in mind that the shear rate dependent
spectrum used above does not provide a basis for the prediction of
any rheological property other than the viscosity. It is not part of a
general theory of nonlinear viscoelasticity but must be defined in
terms of some specific procedure. One might, for example, superpose small amplitude oscillatory shear on a viscometric flow and
determine a shear rate dependent complex modulus, which could,
in turn, be used to calculate H(A, y).
It is of interest here to examine the predictions of Wagner's
equation (3-64) for the viscosity function. The predicted viscosity is:

(4-27)
In order to produce a simple, easily interpretable result, we will use
a discrete relaxation modulus (Equation 2-25) and Wagner's exponential damping function (Equation 3-72). The viscosity is now
given by:

(4-28)
We note that the contributions, T/i( = GiA), of the linear theory are
replaced by the quantity in the brackets, which we have called
T/;( y):

(4-29)
While use has been made of a relaxation spectrum that is entirely
unaffected by the shear rate, the contributions from the terms
involving the largest values of A; decrease much more rapidly than
those for small Ai' as the shear rate is increased. This shows that it
is not necessary to introduce a "truncation" of the spectrum of
relaxation times in order to explain the behavior of the viscosity
function.

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

169

4.5.2 Dependence of Viscosity on Temperature

The dependence of 7]0 on temperature was discussed in Section


2.12. Here we shall be interested in the effect of temperature on the
entire viscosity function, 7]( y). The concept of time-temperature
superposition was introduced in Section 2.12 in connection with the
discussion of the effect of temperature on linear viscoelastic properties. This same concept has been found to be useful as well for
making temperature-independent master curves of viscosity versus
shear rate. In fact, the shift factor for linear viscoelastic properties,
aT' is often found to work equally as well in the shift of the
viscometric functions. Thus, temperature independent representations can be prepared by plotting

7]( y )/a T versus aTy


But if the ratio (ToPo/Tp) is relatively insensitive to temperature,
Equation 2-127 shows that for a given polymer (i.e., a fixed value of
7]o(T aT is proportional to 7]0(T). Thus, a temperature invariant
plot can often be obtained by plotting

7](Y,T)]
.
[ 7]iT) versus [Y7]o(T)]
This procedure was proposed by Vinogradov and Malkin [26,
p. 208] on the basis of the hypothesis that the longest relaxation
time of a polymer governs the viscosity behavior and that this time
is nearly proportional to the zero shear viscosity over the range of
temperature that is of practical interest. Semjonow [28] has proposed the use of the same procedure to prepare a pressure-insensitive, viscosity-shear rate curve, i.e., one plots:

At the end of Chapter 2 we noted that when a quantity does not


involve units of time, no temperature shift factor is necessary. This
implies that a plot of 7] /7]0 versus (J' should be temperature
independent, and this has, indeed, been verified experimentally
[29]. However, because the shear stress becomes somewhat insensitive to changes in shear rate at high shear rates, the curve of

170

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

(71 /710) versus

falls off very sharply at its right end and does not
give a clear picture of the variation of the viscosity.
(T

4.6 THE FIRST NORMAL STRESS DIFFERENCE

As we move down the list of the three viscometric functions, the


experimental difficulties increase sharply at each step. Thus, much
less is known about '\{Ii y) than about 71( y) for polymeric liquids,
and in the case of '\{Ii y) little is known, except that it is a fraction
(between 0.1 and 0.4) of '\{It(y) in magnitude and has the opposite
sign [30].
A convenient point of departure for a discussion of the characteristics of the first normal stress difference is the limiting behavior at
low shear rate. The rubberlike liquid model provides the following
relationship between the limiting low shear rate value of the first
normal stress coefficient and the steady state compliance in the
limit of vanishing stress:

(3-42)
Alternative forms of this relationship are:
(4-30)
(4-31)

Since quite a lot is known about how 710 and IJ depend on


molecular weight and temperature, we can infer the dependence of
'\{It 0 on these variables. Specifically, the steady state compliance is
known to be essentially independent of temperature and molecular
weight for mono disperse polymers, but strongly dependent on the
molecular weight distribution of polydisperse polymers. This observation leads directly to the conclusion that if 710 is proportional to
M a for a monodisperse polymer, then '\{It,O is proportional to M 2a
If a = 3.5, then
'\{I

t,O

ex M

7 .0

(4-32)

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

171

Laun [31] found this dependence to be valid for Nylon-6. With


regard to the molecular weight distribution, we have seen in previous discussions that:
(4-25)

and
'1'10= jOOAH{A)dA
,

(4-34)

Thus, the first normal stress coefficient depends on a higher moment of the relaxation spectrum than the zero shear viscosity. The
spectrum, in turn, is related to the molecular weight distribution,
with a broader distribution leading to a broader spectrum of relaxation times. This means that '1'10 will depend on molecular weight
distribution in a significantly different manner than 7]0' In particular, a high molecular weight tail will have a stronger effect on '1'1,0
than on 7]0'
With regard to temperature dependence, if fP is independent of
temperature, Equation 3-42 implies that:
d log '1'1,0 =
dT

2[ d log

dT

7]0]

(4-35)

The strong temperature dependence of '1'1,0 increases the difficulty of measuring it accurately, especially at the high temperatures
required for many melts. For purposes of comparison, therefore,
either of two resins or of the same resin using two different
instruments, it is advantageous to make use of the temperature
independent property, fP, rather than '1'10'
When M> Me' the first normal stress coefficient is highly dependent on shear rate, and, as in the case of the viscosity, its dependence on molecular weight decreases strongly at high shear rates.
Figure 4-7 shows data for two polymer solutions [32].
The rubberlike liquid model predicts that the first normal stress
difference is quadratic in the shear stress (Equation 4-31), and it
has been observed that in the limit of very small shear rates this is

172

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


105~---------------------------------,

10 4

10 3

10 2
~

'"

(/)

10

Cll

.:;

10

10

10

~----~----~--~----~----~----~

10

10

10

10 2

10 3

10 4

Y (s ')
Figure 4-7. 'l'hi) CUIVes for two solutions: 1: 1.5% polyacrylamide in water and glycerine; 2:
2.9% polybutadiene in Primo!. Data from Ref. 32. Figure adapted from Ref. 2. Copyright
1987 by John Wiley & Sons, Inc. Reprinted by permission.

indeed the case. It has been further observed [33-35] that Nl


continues to be quadratic in (J" for some polymers up to modest
levels of shear stress. If this relationship extends into the region in
which the viscosity follows a power law, we have:
(4-36)
A more general type of power law has also been suggested [34].
( 4-37)
If the viscosity also obeys a power law in the same range of shear

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

173

stress, then:

(4-38)
In terms of the first normal stress coefficient:
'l't a

yan - 2

(4-39)

With regard to time-temperature superposition we note that it


has been found that the time shift factor, aT' which is determined
by shifting viscosity data, often works equally well in shifting normal
stress data. Furthermore, curves of NiaTy) are found to be temperature invariant. This implies that the proper reduced form for
'l't( y) is:

(4-40)
Since aT is nearly proportional to 11o(T) (see Equation 2-126), an
alternative is a plot of:

Finally, as was mentioned at the end of Chapter 2, if all quantities


involving time are eliminated from the relationship, no shift should
be required. Thus, plots of Nt versus (7 are found to be temperature independent [36].
4.7 EMPIRICAL RELATIONSHIPS INVOLVING VISCOMETRIC FUNCTIONS

4.7.1 The Cox-Merz Rules

Cox and Merz [37] observed that curves of 11( y) are often nearly
identical to curves of 111*1 versus w. Thus, the "Cox-Merz rule" can
be expressed as:

(w = y)

(4-41)

This rule has been tested many times and has been found to be
generally reliable for flexible molecules, although there have been

174

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

reports of its failure for some linear [38] and branched [39,40]
polyethylenes. It does not seem to work for rigid molecules [41]. It
is a very useful relationship, because it is easier to determine
11]*(w)1 over a wide range of frequencies than it is to determine
1]( y) over a wide range of shear rates, using only a cone plate
rheometer.
The apparent validity of Equation 4-41 has been a great puzzle to
rheologists ever since it was proposed in 1958. This is because it
implies a universal relation between a linear viscoelastic property
and a nonlinear property. While the linear property is characteristic
of the melt in its equilibrium, fully entangled state, the dependence
of viscosity on shear rate reflects the extent to which shearing has
reduced the entanglement density. Putting it yet another way, we
note that the Wagner equation prediction of the viscosity function
(4-27) shows clearly that 1]( y) depends on both the linear relaxation
modulus function, G(s), and the nonlinear damping function, h( y).
At the same time, the complex viscosity depends only on the
modulus function (see equations 2-61 to 2-66).
Booij et al. [42] have suggested an answer to this puzzle. They
derived the form of the damping function that would yield a
prediction of the viscosity function that is in exact agreement with
the Cox-Merz rule. Their result is:
2
-1
fo(x)dx
y
y

h(y) =

( 4-42)

where fo is Bessel's function of the zeroth order and first kind. This
function oscillates with decreasing amplitude as y increases and is
not at all the sort of damping function that is determined by means
of step shear experiments. However, Booij et al. point out that 1]( y)
is rather insensitive to the exact form of h( y) and that there is
sufficient universality about h( y) for a wide range of polymers that
Wagner's equation predicts a reasonable approximation of the
Cox-Merz rule for a range of possible h( y) functions. This does not
imply that other nonlinear properties are equally insensitive to the
form of the damping function. In fact, time-dependent properties
such as G(t,y) and 1]+(t,y) are quite sensitive to the form of hey).
A second, less used Cox-Merz rule [37] is:

d:~Y)

1]'(w)

(w

y)

(4-43)

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

175

This relates the slope of the viscosity curve to the dynamic viscosity.
There is some evidence that this rule also has wide validity. Booij
et al. [42] have also derived the form of the damping function
implied by Equation 4-43. It is a different function from that given
by Equation 4-42, but again, because of the form of Equation 4-27
this has little effect on the predicted viscosity curve and thus on the
validity of Equation 4-43.
We note that if both the Cox-Merz rules are valid, one can
predict the storage and loss moduli knowing only the viscosity
function.
4.7.2 The Gleissle Mirror Relations

Gleissle [43] has proposed three empirical rules relating viscometric


functions either to each other or to a linear property:

7]+(t)=7]ey)

t=l/y

'l't(t)='I'I(Y)
'l'1(Y)

-2/

00

y/k

(4-44)

t=k/y

(4-45)

x- l [a7](x)]dx
ax

(4-46)

These rules have been evaluated by Friedrich [44] and by Leblans


et al. [45]. The latter authors used Wagner's equation (3-64) as a
basis of their analysis and reached the following conclusions:
1. The damping functions for various polymers are sufficiently

similar to the one required to make Equation 4-44 exactly


valid that it is a reasonable approximation for many materials.
2. 'l'1( y) is more sensitive to h( 'Y) than 7]( y). For this reason, a
material-dependent parameter, k, is required in Equation
4-45. However, Equation 4-45 is not reliable in the range of
shear rates where the viscosity is approaching its constant,
zero-shear value.
3. Equation 4-46 works well, even at low shear rates, although it
has never been tested at substantial shear rates (> 10 s -I)
due to instrument limitations.
Larson [46] and Wissbrun [47] have pointed out that an additional factor that contributes to the validity of the Gleissle relation-

176

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

ships as well as of the Cox-Merz rules, is the breadth of the


molecular weight distribution. In particular, they argue that a broad
molecular weight distribution implies a relaxation modulus that is a
broad function of time. They show that this decreases the sensitivity
of the viscometric functions to the detailed form of the damping
function.
4.7.3 Other Relationships

Laun [48] has proposed the following relationship between 'IIi y )


and the components of the complex modulus:
G' [
G' 2]0.7
'1'1 ( y) = 2 w 2 1 + ( G" )

(4-47a)

An alternative form is:


(4-47b)
He tested this relationship for low and high-density polyethylenes,
polypropylene and polystyrene.
Stastna and DeKee [49] have compared several relationships that
have been proposed to calculate '1'/ y) from 7J( y). These are based
on empirical constitutive equations, and some success in their
application has been claimed. However, Stastna and DeKee conclude that such procedures do not have universal applicability to
polymeric liquids. Like Equation 4-46 they involve integrations of
the viscosity over a range of shear rates that has infinity as its upper
limit, and this information is not available. Furthermore, 'l'l y) is
governed by a higher moment of the relaxation spectrum than 7J( y).
REFERENCES

1. A. S. Lodge, Elastic Liquids, Academic Press, New York, 1964.


2. R. B. Bird, R. C. Armstrong and O. Hassager, Dynamics of Polymeric Liquids,
Vol. 1, John Wiley & Sons, New York, 1987.
3. R. W. Flumerfelt, M. W. Pie rick, S. L. Cooper and R. B. Bird, 1nd. Eng.
Chern. Fundam. 8:354 (1969).
4. S. Middleman, Fundamentals of Polymer Processing, McGraw-Hill, New York
(1977), Chapter 6.

STEADY SIMPLE SHEAR FLOW AND THE VISCOMETRIC FUNCTIONS

177

5. Z. Tadmor and C. G. Gogos, Principles of Polymer Processing, John Wiley &


Sons, N.Y., 1979.
6. S. H. Lin and C. C. Hsu, Ind. Eng. Chem. Fundam. 19:421 (1980).
7. A. C. Dierckes, Jr. and W. R. Schowalter, Ind. Eng. Chem. Fundam. 5:263
(1966).
8. A. V. Shenoy, S. Chattopadhyay, and V. M. Nadkarni, Rheol. Acta 22:90
(1983).
9. D. R. Saini and A. V. Shenoy, Eur. Polym. J. 19:811 (1983).
10. A. V. Shenoy, D. R. Saini and V. M. Nadkarni, Rheol. Acta 22:209 (1983).
11. D. R. Saini and A. V. Shenoy, J. Elastomers Plastics 17:189 (1985).
12. N. G. Kumar, J. Polym. Sci. 15:225 (1980).
13. J. Meissner, KunststoJfe 61:576 (1971).
14. M. M. Cross, in Polymer Systems: Deformation and Flow, Wetton and
Whorlow, editors, Macmillan, London, 1968.
15. F. Bueche and S. W. Harding, J. Polym. Sci. 32:177 (1958).
16. F. Bueche, J. Chem. Phys. 22:1570 (1954).
17. P. J. Carreau, Ph. D. Thesis, Univ. of Wisconsin, 1969.
18. K Y. Yasuda, R. C. Armstrong and R. E. Cohen, Rheol. Acta 20:163 (1981).
19. C. A. Hieber and H. H. Chiang, SPE Tech. Papers 35:1209 (1989).
20. B. Elbirli and M. T. Shaw, J. Rheol. 22:561 (1978).
21. R. A. Stratton, J. Colloid Interface Sci. 22:517 (1966).
22. G. C. Berry and T. G. Fox, Adv. Polym. Sci. 5:261 (1968).
23. C. R. Bartels, B. Crist, L. J. Fetters and W. W. Graessley, Macromolecules
19:785 (1986).
24. M. J. Struglinski and W. W. Graessley, Macromolecules 18:2630 (1985).
25. R. Kumar and Y. P. Khanna, SPE Tech. Papers 35:1675 (1989).
26. G. V. Vinogradov and A. Ya. Malkin, Rheology of Polymers, Mir Publishers,
Moscow, 1980. Also available from Springer-Verlag, Berlin, New York.
27. R. I. Tanner and J. M. Simmons, Chem. Eng. Sci. 22:1803 (1967).
28. V. Semjonow, Adv. Appl. Polym. Sci. 5:387 (1968).
29. G. V. Vinogradov and A. Ya. Malkin, J. Polym. Sci. A2 2:2357 (1964).
30. S. Ramachandran and E. B. Christiansen, J. Non-Newt. Fl. Mech. 13:21
(1983).
31. H. M. Laun, J. Rheol. 30:459 (1986).
32. J. D. Huppler, E. Ashare and L. A. Holmes, Trans. Soc. Rheol. 11:159 (1967).
33. G. V. Vinogradov, A. Ya. Malkin and V. F. Shumsky, Rheol. Acta 9:155
(1970).
34. K Oda, J. L. White and E. S. Clark, Polym. Eng. Sci. 18:25 (1978).
35. R. I. Tanner, Trans. Soc. Rheol. 17:365 (1973).
36. C. D. Han and K-W. Lem, Polym. Eng. Rev. 2:135 (1982).
37. W. P. Cox and E. H. Merz, J. Polym. Sci. 28:619 (1958).
38. L. A. Utracki and R. Gendron, J. Rheol. 28:601 (1984).
39. J. L. White and H. Yamane, Pure Appl. Chem. 28:619 (1958).
40. D. S. Kalika and M. M. Denn, J. Rheol. 31:815 (1987).
41. R. M. Schulken, R. H. Cox and L. A. Minnick, J. Appl. Polym. Sci. 25:1341
(1980).

178

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

42. H. C. Booij, P. Leblans, J. Palmen, and G. Tiemersma-Thoone, J. Polym. Sci.


Polym. Phys. Ed. 21:1703 (1983).
43. W. Gleissle, in Rheology, Vol. 2, edited by G. Astarita, G. Marrucci and
L. Nicolais (Proc. 8th Intern. Congr. Rheol.) Plenum Press, New York,
p. 457 (1980).
44. C. Friedrich, Plaste und Kautschuk 31:12 (1984).
45. P. J. R. Leblans, J. Sampers and H. C. Booij, Rheol. Acta 24:152 (1985).
46. R. G. Larson, Rheol. Acta 24:327 (1985).
47. K. F. Wissbrun, J. Rheo!. 30:1143 (1986).
48. H. M. Laun, J. Rheol. 30:459 (1986).
49. J. Stastna and D. DeKee, J. Rheol. 26:565 (1982).

Chapter 5
Transient Shear Flows Used
to Study Nonlinear
Viscoelasticity
5.1 INTRODUCTION

The flow behavior of molten polymers can be described by the


theory of linear viscoelasticity only when either the total strain or
the maximum strain rate is very small. But there are very few, if
any, commercial melt processing operations in which the flow
satisfies either of these criteria. Thus, while linear viscoelastic
behavior has some application in material characterization, it cannot, in general, be correlated with melt performance in processing
machinery.
In Chapter 3 it was explained that no general theory of nonlinear
viscoelasticity is available. This means that our approach to nonlinear behavior must be mainly empirical. This is a highly unsatisfactory situation, as it means that there is no way of generalizing the
results of one experimental observation so that they help us to
predict how the same material will behave in a deformation different from the one used to make that observation.
To bring some order into the complex world of nonlinear behavior, it is necessary to make use of any concept or technique that
helps organize information about nonlinear behavior. First, we
design our experimental tests in such a way that they are easily
described in terms of one, or at most two parameters. For example,
in steady shear flows, the only parameter is the shear rate. Second,
we present the results of these tests in terms of "material functions"
defined to emphasize a key aspect of the nonlinear behavior. The
aspect usually used as a basis for defining material functions is
179

180

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the extent of the deviation from linear viscoelastic behavior. Thus,


the linear behavior of a material provides a basis for the representation of the nonlinear behavior. Finally, we attempt to describe a
material function in terms of some empirical equation. For example, as explained in Chapter 4, viscosity versus shear rate data can
often be simply described in terms of power law constants; or, as
described in Chapter 3, the results of a series of step shear strain
experiments might be described in terms of an equation for the
damping function, h( y).
In this chapter we will present a number of simply described
deformations that have been used to study the nonlinear viscoelastic behavior of molten polymers. These include, for example, singleand multi-step shear strain, start-up and cessation of steady flow,
and creep and recoil. The material functions normally used to
present experimental results are defined, and example plots of
experimental data are presented. Finally, the use of the damping
function to generalize transient shear flow data is illustrated.
The plastics engineer would like to make use of experimentally
determined nonlinear material functions to solve practical problems. Most often, these problems involve the prediction of processability on the basis of a laboratory test. In this regard it is important
to note that the strain histories used as the basis for the tests
described in this chapter are rarely designed to simulate the flows
that occur in specific plastics forming operations. This is partly due
to the unavailability of rheometers capable of generating complex
deformation patterns. However, it is also a result of the fact that
unless the experiment actually duplicates all essential details of the
process of interest, we have no reliable way to scale up the results
from the laboratory test to the industrial process. Furthermore,
even if we could reproduce the exact kinematics of an industrial
process, we could not extract fundamental resin properties from the
resulting data.
Facing this situation, we generally prefer to measure a welldefined material property that can be related to the chemical and
physical structure of the polymer. In any event, whether or not
there is any correlation between the results of a particular experiment and the performance of a resin in a particular processing
operation must be determined by experience. Of course common
sense suggests that the closer the strain history of the laboratory

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

181

test is to that of the process of interest, the more likely it is that


such a correlation will be found. For this reason, it is important to
have available detailed information about each of the commonly
used test modes, and this is the purpose of the present chapter.
To establish such a processability correlation it is necessary to
measure both the rheological properties and the process performance. Furthermore, it is necessary to compare several polymers
that can meet the product specifications for the process in question.
Few such comparisons have been published, but notable exceptions
are the reports of the IUPAC Working Party on the properties and
performance of LDPE film resins [1,2,3]. The results given in these
reports are discussed in the chapter (no. 17) on film blowing.
5.2 STEP SHEAR STRAIN
5.2.1 Finite Rise Time

In an ideal step strain experiment, a sample is deformed instantaneously, and the strain history is described, mathematically, as a
step function. For a single-step shear deformation this is shown by
curve 1 in Figure 5-1. If the resulting stress is measured as a
function of time, the nonlinear shear stress relaxation modulus can
be calculated as follows:

G(t,y) = IT(t,y)/y

(5-1)

In reality, however, it is not possible to generate an instantaneous


deformation, and the actual strain history will be some function of
time that depends on the mechanical design and the control system
of the rheometer used. A typical strain pattern is shown as curve 2
in Figure 5-1. Obviously there is a finite "rise time."
In order to facilitate the analysis of the effect of a nonzero rise
time, we consider the ramp or constant shear rate pattern shown as
the dashed curve in Figure 5-1. During the starting transient the
strain is given by:

y(t) =

yt

(5-2)

182

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

CD

<i:
a::
.....

'I

'I

'I

"
I

TIME

Figure 5-1. Comparison of ideal and experimentally achievable strain histories for a step
strain test. Curve 1 is the ideal step function; curve 2 is typical of the deformation produced
by a rotational rheometer; dashed line is a ramp approximation of the actual strain history.

and the total strain is:


y

= y~t

(5-3)

where ~ t is the rise time. First, we examine the question of how


large ~t can be and still yield a result that can be interpreted in
terms of a relaxation modulus. Noting that the product IlTfo has
units of time and is closely related to the longest relaxation time of
a melt, Vrentas and Graessley [4] have proposed the following
criterion for linear polymers:

(5-4)
A general rule is that for a rise time of ~t, data should not be
considered meaningful until t > lO~t.
Another approach to this problem is to specify quantitative
techniques for treating data so as to take into account the finite rise
time. For example, Zapas [5] has proposed that the independent
variable, t, of the relaxation modulus be calculated as follows:
t

to -

~t/2

(5-5)

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

183

where to is measured from the instant the shearing begins. Laun [6]
has proposed a more elaborate procedure in which an exponential
damping function (Equation 3-72) is used to calculate a correction
function, f(t o, 'Y, dt), that can be used to determine the true
relaxation modulus, G(to, 'Y), from the measured stress relaxation,
(T(to):

(5-6)

As to becomes much larger than dt the correction function becomes equal to one. Laun points out that if dt is very long the
experiment is more like start-up flow than step strain.
5.2.2 Nonlinear Shear Stress Relaxation Modulus

Single step shear strain is the test most widely used to study
nonlinear viscoelastic behavior. The results of this test are generally
described in terms of the nonlinear relaxation modulus defined in
Equation 5-1. We begin this exposition with a discussion of the
extensive experiments carried out at Kyoto University on concentrated polystyrene solutions.
Figure 5-2 shows the data of Osaki et al. [7] for a solution of a
polystyrene having a very narrow molecular weight distribution. The
molecular weight is 8.43 X 10 6 , and the concentration is 0.06 g/cm 3
The strain magnitude ranged up to 6.1. For strains less than 0.57,
the relaxation modulus was found to be independent of strain, and
this was taken to be the linear modulus and is the uppermost curve
in Figure 5-2. The data start near the end of the plateau zone and
extend well into the terminal zone. As the strain is increased, the
curves deviate from the linear behavior and fall successively further
below it. At the largest strains, the shape becomes more complex,
with two inflection points. This suggests the appearance of a new
relaxation process that does not manifest itself in the linear behavior. This behavior is similar to the prediction of the Doi-Edwards
theory, which is shown in Figure 3-10. According to this theory, the
new relaxation process that causes the change in shape is retraction
within the tube (contour length relaxation). The characteristic time
for this process is the longest Rouse relaxation time, AR , which is
between Ae and Ad.

184

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

a.'" 10'
,..:;

;>-

...:;

(,!)

10

10'

10 2
TIME,s

Figure 5-2. Relaxation modulus for a polystyrene solution in chlorinated biphenyl (M = 8.42
x 10 6 , C = .06g/cm 3 ) at 30C for several strain magnitudes. Strains are, from top to bottom:
< .57 Oinear behavior), 2.06, 4.0 and 6.1. Adapted from Ref. 7. Copyright 1982 by The
American Chemical Society. Reprinted by permission.

The theory predicts that when the time is greater than the
retraction time (A R ), the nonlinear stress relaxation becomes separable, i.e., that it can be expressed as the product of the linear
modulus and a damping function, as shown by Equation 3-68.
G(t,y)

h(y)G(t)

(3-68)

To test this, Osaki et al. [7] made a plot of G(t,y)/h(y). This is


shown in Figure 5-3. This quantity should be the linear relaxation
modulus, G(t), which is independent of y. This is indeed what was
observed at times greater than a time Ak , which was found to be
independent of y and approximately proportional to M2.
The values of the damping function resulting from the vertical
shift of the data are shown in Figure 5-4, together with the
prediction of the Doi-Edwards theory. There is qualitative agreement, although all of the data fall somewhat above the theoretical
curve.
Separability fails at times less than Ak , where the experimental
curves diverge sharply and become strongly dependent on y. The
Doi-Edwards theory explains this as a manifestation of retraction or

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

185

103~~--------------------------~

co

Il..

10 2

'E'

-<5""
~

10'

10'
TIME,s

Figure 5-3. Nonlinear relaxation modulus divided by the damping function, as a function of
time, for the polystyrene solution of Figure 5-2. Adapted from Ref. 7. Copyright 1982 by
The American Chemical Society. Reprinted by permission.

10-2~------~------~~~-J

10-'

10

10'

Figure 5-4. Damping function determined from the vertical shift of the long-time portion of
the G(t,"Y) curves shown in Figure 5-2. The dashed curve shows the prediction of the
Doi-Edwards theory. Adapted from Ref. 7. Copyright 1982 by The American Chemical
Society. Reprinted by permission.

186

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

contour length relaxation, a relaxation mechanism that has negligible effect on linear behavior but becomes important for large, rapid
deformations.
Osaki et al. [8] have measured the nonlinear stress relaxation of
solutions of blends of two polystyrenes having different molecular
weights.
The data described above and shown in Figures 5-2 to 5-4 are for
a solution in which eM (e = concentration in g/cm 3 ) was 5 X 10 5
The behavior shown has been called "Type I" by Osaki and Kurata
[9] in order to distinguish it from the behavior of solutions with very
high values of eM. This second type of behavior, which they called
"Type II," is illustrated in Figure 5-5 for a solution of polystyrene
in diethylphthalate with e = 0.221 g/cm 3 and M = 5.53 X 10 6 [10].

Figure 5-5. Relaxation modulus of a Type II polystyrene solution. (M = 5.53 X 10 6 , C = .22


g/cm3, Tref = 30C) Strains are, from top to bottom: < .22 (linear behavior), .44, 1.11, 3.34,
6.68, 10. Adapted from Ref. 10. Copyright 1975 by John Wiley & Sons, Inc. Reprinted by
permission.

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

187

10'

t,

Figure 5-6. Data of Figure 5-5 shifted vertically to bring curves together for t> AI' Adapted
from Ref. 9. Copyright 1980 by The American Chemical Society. Reprinted by permission.

We see in Figure 5-6 that superposition is now possible only at


times well into the terminal zone (t > AI)' At all shorter times the
behavior is highly irregular. Osaki and Kurata [9] found in their
studies that Type I behavior was observed when eM < 10 6 , while
Type II behavior was observed at higher values of eM. This critical
value of eM corresponds to about 50 entanglement points per
molecule (MjM). Vrentas and Graessley [11] found that the
critical value of eM for polybutadiene was about 150,000. This
suggests that a general criterion for the occurrence of Type I
behavior is that the ratio M j Me be less than about 30.
The failure of separability for very high molecular weight polymers is not understood at the present time. It is counter to the
prediction of reptation theories whose validity should be enhanced
by an increase in molecular weight.

188

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

5.2.3 Time-Temperature Superposition

Laun [6] found that his relaxation modulus data for low density
polyethylene "Melt I" were separable, i.e., that they could be
described in terms of a damping function, h( y), multiplied by the
linear relaxation modulus, G(t), as indicated by Equation 3-68. He
further noted that the damping function was independent of temperature, implying that a time-temperature superposition principle
suitable for use with the linear relaxation modulus can also be used
to bring together G(t, y) data for several temperatures.
Einaga et al. [12] and Fukada et al. [10] found that their relaxation modulus data for solutions of narrow MWD polystyrene were
subject to time-temperature superposition at a given value of the
strain magnitude, even when separability was not observed. Their
superposition procedure consisted of plotting [G(t, Y)PoTolpT] versus [tlaT]' where aT is the horizontal shift factor determined from
linear viscoelasticity studies. The success of this superposition procedure implies that the relaxation times do not depend on the
strain, even when Equation 3-68 is not valid.
5.2.4 Strain-Dependent Spectrum and Maxwell Parameters

Where separability fails, relaxation modulus data can be interpreted in terms of a strain-dependent relaxation spectrum, H(A, y)
or a nonlinear form of the generalized Maxwell model:
G(t,y) =

~GJy)exp[ Ai~;) 1

(5-7)

An alternative form is:

(5-8)
Fukada et al. [10] have shown plots of H(A, y), G i( y) and Ai( y) for
Type II polystyrene solutions. The strain-dependent Maxwell parameters were determined by means of a graphical procedure [13].
Figure 5-7 shows plots of G1(y), Giy), A1(y) and Aiy) for the

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

189

G, (Pa)

Ai (S)

10
SHEAR STRAIN, y

Figure 5-7. First two moduli (G) and relaxation times (,\.) of Equation 5-7 for the
polystyrene solution of Fig. 5-5. Adapted from Ref. 10. Copyright 1975 by John Wiley &
Sons, Inc. Reprinted by permission.

same polymer solution used to obtain Figures 5-5 and 5-6 [10]. The
two longest relaxation times are independent of strain.
To describe phenomena governed primarily by the longest relaxation times, Osaki et al. [14] note that if these values of Ai are
independent of strain, the relaxation modulus can be expressed as
follows:
G(t,y)

Ehi(y)Gie- I /

A;.

(5-9)

;=1

The products, [h i( y)G;J are termed "relaxation strengths." This


form could be used for Type I materials when t < tk and for Type
II for t < AI' Osaki et al. [14] found that for Type I materials, hl(y)
and hi y) could be approximated satisfactorily as simple exponentials, whereas for Type II materials a more complex function form
was required.
Luo and Tanner [15] used the concept of a series of damping
functions corresponding to the various relaxation times to formulate a nonseparable constitutive equation having the form of Equa-

190

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

tion 3-62, with the nonlinear memory function given by:


(5-10)
The damping functions were assumed to be of the form of Equation
3-81 but with a different value of f3 for each relaxation time:
h=
I

----------~----------~

(1 - 3a) + a[f3Jl + (1 - f3;)I2 ]

(5-11)

5.2.5 Normal Stress Differences for Single Step Shear Strain

In Chapter 3 it was demonstrated that the rubberlike liquid model,


Wagner's equation, and the Doi-Edwards theory predict that in a
step strain experiment:
(3-70)
or:

This is called the Lodge-Meissner relationship, and it has been


found to be valid for a branched polyethylene "Melt I" [6],
polystyrene solutions [4,16,17], molten polystyrene [18] and for
polybutadienes [4,11,19].
The rubberlike liquid model and Wagner's equation both predict
that the second normal stress difference is zero, but the DoiEdwards theory predicts nonzero values of Nit, y)/Nit, "I). This
latter ratio is predicted to be negative and independent of time and
the same function of "I for all linear, monodisperse, entangled
polymers. Whereas the independent alignment assumption has little
effect on the prediction of shear stress and the ratio Nit)/a(t), it
has a significant effect on the normal stress ratio. For example, in
the limit as "I approaches zero, the general theory predicts that

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

191

-1/7, while with the independent alignment assumption


the value becomes - 2/7.
Osaki et al. [17] measured N21 Nl for a polystyrene solution and
found the results to be in qualitative agreement with the prediction
of the Doi-Edwards theory for strain magnitudes up to 4.0. At low
values of 'Y, N21 Nl approached a constant value of about - 0.2,
while for 'Y > 2 its absolute value decreased rapidly.
Kimura et al. [19] also measured N21Nl for a branched, broad
distribution polybutadiene. They found this ratio to be independent
of time and fitted their data to the following equation for
N21Nl

0.5

< 'Y < 3.5:

(5-12)

5.2.6 Multistep Strain Tests

Multistep strain experiments have been used to evaluate certain


features of constitutive equations. For a material that obeys the
Boltzmann superposition principle (Equation 2-10) the measured
response to a single step strain deformation, i.e., G(t), can be used
to calculate the response of that material to any sequence of step
strains in any direction. For example, the response to the double
step strain shown in Figure 5-8 would be as follows, for t > 0:

(5-13)

r
I

- t,

o
TIME,

Figure 5-8. Double step strain test.

192

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

However, for materials behaving in a nonlinear way such a


superposition is no longer valid, and a nonlinear theory is required
to predict the stress. For example, according to Wagner's equation
(3-64) the response to the double step shear strain would be as
follows, for t > 0:
a(t) = CYI

+ Y2)h(YI + Y2)G(t + t I )

+ Y2 h ( Y2)[ G(t)

- G(t

+ t I )]

(5-14)

Vrentas and Graessley [4] and Osaki et al. [20,21] used the double
step strain to study the nonlinear behavior of several polymer
solutions. They found that the response to the second strain could
be predicted accurately by use of the Doi-Edwards theory, with or
without the independent alignment assumption, i.e., Equation 5-14
was found to be valid, as long as the second strain was in the same
direction as the first.

A more severe test of theories of nonlinear viscoelasticity is the


strain history shown in Figure 5-9, in which the second step strain is
in the opposite direction from the first. It was mentioned in Section
3.7.5.1 that this test has been used to evaluate the validity of the
"irreversibility assumption" as a modification of Wagner's equation.
Figure 3-9 shows that this modification is necessary to avoid a
substantial overprediction of the shear stress for the case of LDPE.

Z
<C
a:

Y,
-Y2

en 0

- 1,

0
TIME,I

Figure 5-9. Double step strain with reversal.

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

y,

193

Y3

-t,

0
TIME,

Figure 5-10. Spike strain test.

Figure 3-9 also shows that when 11 is larger than 12 and t1 is short,
the shear stress can change sign during the relaxation process for
t> O.
When t1 is large, and 11 and 12 are not too large, both the shear
stress and the first normal stress difference can be predicted by the
Doi-Edwards theory, with or without the independent alignment
assumption [20], but for the more severe case of short t 1 and large
strains, the independent alignment assumption is no longer valid,
especially in the case of the shear stress [4,19,20]. For Type II
materials, of course, the theory is of no use except at quite large
values of t.
Another multistep shear strain that has been used to evaluate
constitutive equations is the "spike strain test" [22] shown in Figure
5-10. We note that according to the Boltzmann superposition principle, the spike at time - t 1 has no effect on the response to the
subsequent strain at time t = O.

(t> 0)

(5-15)

The rubberlike liquid model also predicts this stress, while Wagner's

194

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

equation (3-64) predicts the following when

> 0:
(5-16)

Unless 'Yl is very small, however, the irreversibility assumption does


predict an effect of the spike on the subsequent stress relaxation at
t = o. Bruker [23] has used a similar, triple step shear strain
experiment to evaluate constitutive assumptions.
5.3 FLOWS INVOLVING STEADY SIMPLE SHEAR

It was noted in Section 5.2.1 that it is impossible to generate an

ideal step strain deformation; this would only be possible with an


infinite velocity and an inertialess (massless) test material. Furthermore, it is even difficult to produce an acceptable approximation of
a step shear strain, because the motion control systems of many
rheometers are not well suited to this mode of operation.
By contrast, the start-up and cessation of steady shear and the
sudden change of shear rate are relatively easy to generate in the
laboratory. It is the velocity rather than the position of the moving
surface that must be controlled, and this velocity is always finite.
For this reason, the most popular method for studying nonlinear
viscoelasticity is to track the stresses as functions of time in flows
involving steady simple shear. In this section, the standard test
modes are described, and the relevant material functions are
defined.
5.3.1 Start-up Flow

In this test, a sample initially in its equilibrium state is subjected to


a constant shear rate, y, starting at time t = O. The rheologically
significant measurable quantities are the shear stress, a, and the
first and second normal stress differences Nl and N 2 . The relevant
material functions are defined as follows: Shear stress growth
function:
(5-17)

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

195

Shear stress growth coefficient:


(5-18)
First normal stress growth function:
(5-19)
First normal stress growth coefficient:
(5-20)
Second normal stress growth function and coefficient:

Nt(t,1') == 0"22(t,1') - 0"33(t,1')

(5-21)

'l'2+(t,1') == Nt /1'2

(5-22)

In the limit of small shear rate, 1] +(t, 1') will become equal to the
corresponding material function of linear viscoelasticity:
(5-23)
At sufficiently long times, the stresses will become steady, and the
values of the material functions will approach the corresponding
viscometric functions. For example,
(5-24)
(5-25)
lim [Nt(t, 1')]

t->oo

Nz{ 1')

(5-26)

Figures 5-11 and 5-12 show the curves for 1]+(t,1') and 'l't(t,1')
for a solution of polystyrene in tri-cresyl phosphate [24]. The
dashed curve is calculated from the linear spectrum using Equation
2-85. As the shear rate is increased, the 1] + curves deviate from the
one for linear behavior at progressively shorter times. Also they

196

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

y=

.107

~---==.170

.-::::;...-----.427
;~-----1.07

(/J

til

11.

- - - - - - - 2.69

~ 10 2
...:

- - - - - - - 6.77

T~

"-----17.0
' - _ - - - - 42.7

V_ _ _ _ _L-_ _ _ _~==~======~1~0~7_ _L_ _~


10 1.0
10.0
0.1
TIME, S

Figure 5-11. Shear stress growth function for a polystyrene solution. Dashed line is calculated from the linear spectrum using Equation 2-85. Adapted from Ref. 24. Copyright 1977
by Steinkopff Verlag. Reprinted by permission.

exhibit a stress overshoot, with the maximum shear stress occurring


at a time ts and the maximum value of Nt occurring at time tN'
Over the lowest decade of shear rates, these times are inversely
proportional to the shear rate, implying that the maxima always
occur at constant values of the shear strains, 'Ys and 'YN' in this
range of shear rates. The ratio, tN/t s , (equal to 'YN/'YS) is between
2 and 2.3 over the entire range of shear rates studied.
Osaki et al. [14] used Equation 5-9 to fit their stress relaxation
data for polystyrene solutions and then determined the corresponding memory function and used Equation 5-27 to predict the shear
stress growth coefficient.

(5-27)
They found that the result was not very sensitive to the exact form
of the damping function. Using a damping function having the form
of a simple exponential (Equation 3-72 with n = 0.37) for a Type II

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

197

10 2

'"

'"
I

0.

~ 10

---....1.----...1...-.....
1.0
10

0.1 .....- -.......

0.1

TIME.

Figure 5-12. First normal stress growth function for a polystyrene solution. Adapted from
Ref. 24. Copyright 1977 by Steinkopff Verlag. Reprinted by permission.

solution they computed that:


Ys

YN

n
2

(5-28)
(5-29)

Thus:
Ys

(5-30)

These results were in fairly good agreement with the experimentally


determined values of these quantities. We note that Equation 5-30

198

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

is in approximate agreement with the observations of Graessley


et al. [24] for a different polystyrene solution.
Wagner [25] gives the following equations for calculating the
stress growth functions on the basis of a simple exponential damping function:

7]+(t,y)

[G(s)[exp( -nys)](l - nys) ds


o

(5-31)

'l't(t, y)

[G(s)s[exp( -nys)](2 - nys) ds


o

(5-32)

Laun [6] gives the integrated forms of Equations 5-31 and 5-32
when G(s) is represented by a set of Maxwell elements
(Equation 2-25).
Wagner [25] used the above equations to analyze Meissner's data
[26] for LDPE "IUPAC A." Wagner found that n = 0.143 gave a
good fit of Meissner's data, and this implies that:
'Ys::::: 7
'YN:::::

14

These are somewhat different from the values of Graessley et al.


[24] for a polystyrene solution.
Laun [6] used the damping function to analyze his data for LDPE
"Melt I." He used both a single exponential h( 'Y), Equation 3-72,
and a summation of two exponentials, Equation 3-73, and compared the predictions of the two. 1 The results are shown in Figure
3-4. Laun found that the general shape of the predicted curve and
the maximum values of a+(t, y) and Nt(t, y) were rather insensitive to the functional form of the damping function and that the
position of the maxima, i.e., the values of t s and tN' were the most
sensitive features of the curves. The experimental curves yielded a
value of tNlts of about 3. The single exponential damping function
predicted a value of 2, while the double exponential form predicted
IThe parameters of the two damping functions for this resin are given in Section 3.7.4.1 of
this book.

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

199

a value of about 3. The precise values of tN and ts are often


difficult to determine precisely because the maximum in the stress
is broad and flat.
Menezes and Graessley [27] determined '17 +(t, y) and Nt(t, y)
for several solutions of polybutadiene having narrow molecular
weight distributions. The departure of the curves from those in the
range of validity of the rubberlike liquid model always occurred at a
shear strain of about 0.9, for both '17 + and Nt. At low shear rates,
'Ys was about 2.0 and 'YN was about 5.0. At higher shear rates, these
quantities increased with shear rate, although the ratio 'YNI'Ys was
always between 2.1 and 2.5. Menezes and Graessley discussed the
functions 'YN(Y) and 'Yiy) in terms of the Doi-Edwards theory.
A stress "undershoot" following the initial overshoot was observed in early experiments and is predicted by some theories [27],
although these are not always observed [14,27].
Meissner [28] has reported the results of an interesting experiment in which polyisobutylene was subjected to start-up shear flow
at time t = 0, and after a time, te , steady simple shear motion in the
X3 direction was superposed. For 0 < t < t e , the T21 component of
the stress followed the expected pattern of the stress growth function, u+(t), and T23 was zero. The introduction of the additional
shearing in the X3 direction altered the behavior of T21(t) and
generated a shear stress component, T2it). These are shown in
Figure 5-13 for three values of t e, with Y23 = 0.1 S-1 for t > te' The
introduction of Y23 at time te accelerates the approach of T21(t) to
its steady state value. In addition, the overshoot in T 23 is substantially reduced by the preshearing in the x 1 direction.
5.3.2 Cessation of Steady Shear Flow

In this test, a fluid is subjected to steady simple shear until all the
rheologically meaningful stresses are steady. At time t = 0, the
motion is suddenly halted, and these stresses are measured as
functions of time. The relevant material functions are as follows:
Shear stress decay function:
(5-33)

200

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1
1

10.6

31

50

1001

150

200

104.3
TIME, S

Figure 5-13. Meissner's results for bidirectional start-up of simple shear for polyisobutylene
at 25C. Shear in x I direction is started at t = 0; at times indicated by dashed lines, shear in
x3 direction is started. Curves starting at t = 0 show values of 721' others show 723' The shear
rate is 0.1 s -1. Adapted from Ref. 28. Reproduced with permission from the Annual Review
of Fluid Mechanics, Vol. 17, Copyright 1985 by Annual Reviews Inc.

Shear stress decay coefficient:


(5-34)

First normal stress decay function:


(5-35)

First normal stress decay coefficient:


(5-36)

Second normal stress decay function and coefficient:

N;(t,,,;)

== 0'22(t,,,;)

'l'2-(t,,,;)

==--+

N'Y

- 0'33(t,,,;)

(5-37)
(5-38)

TRANSIENT SHEAR FLOWS USED TO STUDY NONUNEAR VISCOELASTICITY

1.0

201

r--------------,

........

........

....... .......
0.85 ....

.01

10.8

21.4

TIME,s

Figure 5-14. Shear stress decay coefficient for a solution of polyisobutylene. Dashed line is
calculated from the linear spectrum using Equation 2-89. Adapted from Ref. 27. Copyright
1982 by John Wiley & Sons, Inc. Reprinted by permission.

The initial values of these functions (t = 0) are obviously equal to


the corresponding viscometric function at the same shear rate. In
the limit of very small shear rates, '11 - (t, y) becomes equal to the
corresponding function of linear viscoelasticity '11- (t), while 'It1- (t, y)
is thought to approach the function predicted by the rubberlike
liquid model.
Figures 5-14 and 5-15 show '11-(t, y) and 'lt1-(t, y) curves measured by Menezes and Graessley [27] for a polyisobutylene solution.
As the shear rate increases, the relaxation of the stresses occurs
more rapidly, showing that preshearing affects the relaxation
process.
Osaki et al. [14] have shown that for a material whose memory
function for shear deformations is given by:

M(s,y)

= ~hj(Y)ie-S/A;
I

(5-39)

the shear stress decay coefficient is:

'11-(t,y)

Egj(Y)Ajexp( -tIA;)
j

(5-40)

202

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1.0

0.3

.01 ......- - - ' - - - - - - ' - - - - - '

TIME, S

Figure 5-15. First normal stress decay coefficient for the polyisobutylene solution of Fig.
5-14. Dashed line is calculated from the rubberlike liquid model using Equation 3-45.
Adapted from Ref. 27. Copyright 1982 by John Wiley & Sons, Inc. Reprinted by
permission.

where the "relaxation strengths," g;( y), are given by:

g;(y)

GfOOh;(ys)exp( -sIA;}sds
---iA;

(5-41)

For a separable memory function with a single exponential damping function, Laun [6] gives the following equations:

- .-

rJ(t,y)-L.J

[G;A; exp( -tIA;} 1

;=1

- .-

'1(t,y)-2L.J

;=1

.2

(1 + nyA;)

[G;A7 exp( .-tIAJ 1


(1 + nyA;)

(5-42)
(5-43)

According to these expressions, the relaxation times are unaffected


by the preshearing, but the relaxation strengths, which are G; in the
linear theory, are replaced by the following, shear-rate dependent

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

203

quantities:
(5-44)
Laun found that using the value of n determined from step strain
experiments for LDPE "Melt I," Equations 5-42 and 5-43 were in
good agreement with experimental results, and that the agreement
was rather insensitive to the specific functional form of the damping
function. This phenomenon has also been noted by Osaki et al. [14]
in their studies of polystyrene solutions.
Because this experiment involves only relaxation processes and
does not track the response of the sample to a large deformation,
the decay functions do not have as many interesting features as the
stress growth functions. Thus, for purposes of empirical characterization of nonlinear behavior, start-up flow is considered more
useful than cessation of steady shear.
5.3.3 Interrupted Shear

Relaxation processes can only be monitored experimentally as long


as the stresses are sufficiently large to be measured with some
precision. However, it has been hypothesized [29] that a "reentanglement" process may continue to occur long after the stresses have
fallen to insignificant levels. A method of tracking the long-term
approach to equilibrium is the interrupted shear experiment [29, 30].
The strain history and typical shear stress pattern for this test are
shown in Figure 5-16. The sample is initially subjected to start-up
flow. After the stresses reach steady values, shearing is stopped for
a period of time, fro called the "rest time," and the flow is restarted
at the same shear rate as before. At high shear rates, the shear
stress will have a maximum value, am' during the first start-up cycle.
In subsequent cycles, the value of am will depend on the rest time.
Thus, even when fr is sufficiently long that the stress has decayed to
levels too low to measure, the state of the sample can be determined from its response to the next shearing cycle. Thus, the peak
stress is a function of the rest time as well as the shear rate.

204

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

a::w Y
Iw
Via:: 0

~t'-1

en
en
w
a::
Ien
a::
w
:r:
en
0
0

TIME

Figure 5-16. Strain history and typical stress response for interrupted shear.

For sufficiently long rest times, (Tm will approach a value characteristic of a melt that is initially in its equilibrium state, (Tm( 00, y).
Figure 5-17 shows a plot of (Tm(t) for a high density polyethylene at
1700 e [31] at a shear rate of 0.10 S-1. Even at this low shear rate
the viscosity curve is markedly non-Newtonian, and strongly nonlinear behavior is to be expected.
Stratton and Butcher [29] have suggested the use of an empirically defined "reentanglement" time, te' to characterize the response of a polymeric liquid in this test. It is the time required for
the peak stress minus the steady state stress (r6) to recover to a
level within lie of the initial peak (In(e) = 1). Thus:
(5-45)

For polymer solutions [29] and melts [31] this quantity has been
found to be significantly longer than the time required for 11 - Ct, y)
to fall to a value of 11( Y)Ie. Tsang and Dealy [32] have compared
these observations with the predictions of several constitutive equations. They have shown that the large value of te arises in some
models from the fact that the expression for (TmCtr) involves a

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

205

5.2

5.0

eo

4.8

rr-

4.6

l-

Q.

""

oE

I-

4.4

lI-

4.2
10

10 2
REST TIME,

Figure 5-17. Peak stress as a function of rest time for interrupted shear of a HDPE at 170e.
The shear rate is 0.1 S-I. Adpated from Ref. 31. Copyright 1981 by John Wiley & Sons,
Inc. Reprinted by permission.

different weighting of the relaxation times than do those for G(t, y)


or TJ -(t, 1)
The interrupted shear test provides information about very longterm responses to shearing that cannot be followed by measuring
stress relaxation. It is thus related to the phenomenon of "shear
modification" in which shearing at a high rate can alter the response of a melt to a subsequent melt forming operation [33-36].

5.3.4 Reduction in Shear Rate

The experimental determination of the reentanglement time is a


time consuming process. A related test that provides information
about the time scale of the reentanglement process from a single
test is one in which the shear rate is suddenly reduced from 11 to
12 at time t = O. This strain history is illustrated in Figure 5-18. An
empirically defined reentanglement time for this test, t E> is related
to the time required for the stress to recover from its minimum

206

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

y,

a:UJ
IUJ
Via:

Y2
0

CfJ
CfJ

UJ
a:
ICfJ
a:

UJ
:r:

CfJ

TIME

Figure 5-18. Reduction in shear rate experiment.

value,

<Tmin ,

to within 1je of its steady state value, TI( Y2))'2 [31].

(5-46)
An alternative definition can be formulated by replacing <Tmin by
TI( Ylh2
Immediately after the reduction in shear rate the material continues to behave like a material having the viscosity and relaxation
spectrum corresponding to the initial shear rate, YI. Only over a
period of time does its viscosity recover to its steady state value
corresponding to the second shear rate. Dealy and Tsang [31] have
reported values of tE for high and low density polyethylenes.
5.4 NONLINEAR CREEP

In the tests described above, the shear rate is the parameter that
describes the strain history. In the case of creep it is the stress that
is held constant while the strain is measured. The shear creep

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

207

compliance is defined as follows:

y(t)

(5-47)

J(t,u)=u

At sufficiently long times the strain rate will become constant and
equal to the stress divided by the viscosity.
lim

[d Y ] =

dt

t-->oo

(5-48)

17

Thus, the curve of the compliance versus time ultimately becomes a


straight line with slope 1/17. The t = 0 intercept on the J(t) curve
is called the steady state compliance, Js(u), and the equation of the
line is:
J(t,u)

+-

Js(u)

17

(5-49)

where the viscosity corresponds to the shear stress, u.


For sufficiently small levels of stress, linear viscoelastic behavior
will be exhibited, and we have:
lim [J ( t , u )]

u-->O

J(t )

(5-50)

and the equation of the long-time straight line portion of the curve
is:
J(t) =J2

+170

(5-51)

where J2 is the steady state compliance for linear viscoelastic


behavior.
The general characteristics of the creep compliance curve are
illustrated in Figure 5-19. Curves are shown for three stress levels,
u 1 , u 2 , and u 3 , where U3 > U 2 > U 1 and U 1 is sufficiently small that
linear viscoelastic behavior is exhibited. The main features of these

208

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

"

:::;-

JO
5

:::i

a.. Js ( (2)
~

0
0

a.. Js (03)
w
w
a:

o~------------------------------------~
o
TIME

Figure 5-19. Sketch showing the general characteristics of the curve of creep compliance
versus time at three stress levels. 0"3 > 0"2 > 0"1; at the lowest stress linear behavior is
exhibited.

curves are as follows:


1. The level of the compliance increases as the stress is lllcreased.
2. The steady state compliance decreases as the stress is lllcreased.
3. The time to reach steady state decreases as the stress increases.
4. At short times the creep compliance always follows the linear
viscoelastic curve. The departure from the linear curve occurs
at a certain value of the strain. Thus, the time at which the
departure occurs decreases as the stress is increased.
It has been found [37,38] that the steady state compliance can
still be dependent on stress at shear rates well within the range
where the viscosity is independent of shear rate. This observation
shows that one cannot conclude on the basis of a single material
function that a material is sufficiently close to its equilibrium state

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

209

3.0
";"

0.4

<U

a.

-"

6 2.0

::;-

0.2
1.0
0

TIME, S

Figure 5-20. Creep compliance data for a polystyrene solution. The lowest curve shown is
within the linear regime. The inset has expanded scales to show the short-time behavior.
Adapted from Ref. 39. Copyright 1984 by John Wiley & Sons, Inc. Reprinted by
permission.

that all aspects of its rheological behavior will be in accord with the
Boltzmann superposition principle.
Figure 5-20 shows creep compliance curves reported by
Nakamura et al. [39] for a polystyrene solution at several stress
levels. The deviation from linear behavior (the lowest curve shown)
was found to occur at a critical strain that depended on the solution
concentration.

5.4.1 Time-Temperature Superposition of Creep Data

Wagner and Laun [40] measured both the strain, y(t, u, T) and the
first normal stress difference, Nit, u, T) for the creep of LDPE
"Melt I" at several temperatures. They found that both quantities
could be plotted such that the curves were temperature indepen-

210

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

dent by use of a time shift factor, aT. In other words, it was found
that
(5-52)

and
(5-53)

where To is a reference temperature. The shift factor for both


quantities was found to agree with that obtained by shifting viscosity data. This implies that:
(2-127)

Furthermore, the viscosity had an Arrhenius temperature dependence (Equation 2-128).


5.5 RECOIL AND RECOVERABLE SHEAR

If, during any shearing deformation, the shear stress is suddenly


reduced to zero, a viscoelastic liquid will spring back in a direction
opposite to that of the original shearing deformation. Flow in the
x 2 direction is not permitted, and this phenomenon is thus called
"constrained recoil" or "constrained recovery." The recoverable
shear, 'Yoo' is one measure of the level of molecular orientation that
existed at the instant the shear stress was removed [41]. It is not a
unique measure, however, and different measures, such as birefringence, are also used.
5.5.1 Creep Recovery

When the recoil results from the interruption of a creep experiment


at a time to after the sample was first loaded, it is called "creep
recovery." A sketch showing how the strain varies with time in such
an experiment is shown in Figure 5-21. The recoil strain, 'Yr> is

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

211

------yoc

<C

II:
I-

en

II:
<{

W
I

en

to
TIME

Figure 521. Sketch showing essential features of a creep recovery curve.

defined so that it is a positive quantity:


(5-54)
If to is sufficiently large that the creep portion of the test has

reached its steady state regime, the recoil becomes independent of


to, and we let t represent the time since the shear stress was

eliminated. This is a classical creep recovery test, and the relevant


material functions are defined as follows: Recoil strain

,At,a) == y(O) - y(t)

(5-55)

Recoil function (Recoverable compliance):

R(t,a) == y,
a

(5-56)

Ultimate recoil (Recoverable shear):

'Yc,,( a) == lim [ y,( t, a)]


/-+00

(5-57)

Ultimate recoil function:

Kx>( a) == lim [R( t, a)]


/ -+00

(5-58)

212

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

:::"""' -3

Ol

.2 -4

109(1, s)

Figure 5-22. Recoil function of LDPE "IUPAC Coo at l30C at three stress levels; in Pa,
from top to bottom: 6.0, 360, 1469. Adapted from Ref. 37. Copyright (\) 1977 by John Wiley &
Sons, Inc. Reprinted by permission.

Since a = 7]Y at t = 0, the shear rate can be used in place of the


stress as the parameter of these material functions.
For sufficiently low levels of stress, linear viscoelastic behavior
will be exhibited, and the material functions become independent
of a (or Y)'
lim [R(t,a)]

R(t) =J(t)-

IT-->()

lim [R"Ja)] == R~

IT-->()

J~)

7]0

(5-59)
(5-60)

Figure 5-22 shows the recoil function of LOPE "IUPAC C"


corresponding to several levels of the creep stress [37]. We note
that the recoil function and the ultimate recoil function decrease
with increasing creep stress. Also, the approach to the steady state
value occurs more rapidly as the creep stress increases.
The three IUPAC resins, A, B, C, had been found to he almost
indistinguishable on the basis of the commonly measured rheological properties, but they hehaved quite differently when used to
make blown film [1]. Agarwal and Plazek [37] found that the
behavior of the recoil function at long times and moderate creep
stress revealed differences between the three resins, as shown in
Figure 5-23. However, it is not possible to say that there is a
connection between the processibility behavior and the recoil ohservations.

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

213

--2~------------------------------------'
I

~ -3

-4

109(t, 5)

Figure 5-23. Recoil functions for IUPAC LDPE resins A, B, and C [1] at 130a C and a creep
stress of 6.0 Pa. Adapted from Ref. 37. Copyright 1977 by John Wiley & Sons, Inc.
Reprinted by permission.

5.5.1.1 Time-Temperature Superposition: Creep Recovery

Wagner and Laun [40] reported the recoil strain and first normal
stress difference following the steady shear rate stage of a creep test
for LDPE "Melt I" at several temperatures. They found that both
quantities could be plotted in such a way that the curves were
temperature invariant by use of a time shift factor, aT' which was
the same as that obtained by the shift of viscosity data. In other
words, it was found that

(5-61)
and

(5-62)
where To is the reference temperature.
In his measurements of the ultimate recoil, 1'00' corresponding to
creep time, t, for the same polymer, Meissner [42] found that a
temperature invariant curve could be obtained by using the creep
strain, 1'([), as the independent variable in place of time.

(5-63)

214

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

' ......

f/)
f/)

W
II:
l-

---

----------

f/)

ot--_.1
z

II:
lf/)

ot---"'
o

to
TIME

Figure 5-24. Sketch of stress and strain CUlVes for recoverable shear following start-up of
steady simple shear. Dashed line shows how stress develops if recoil is not allowed.

5.5.2 Recoil During Start-up Flow

Constrained recoil experiments provide information about the


stored elastic energy and extent of molecular orientation during any
simple shearing deformation, yet). For example, if a start-up flow is
suddenly interrupted at time to, by the removal of the shear stress,
the recoverable shear Yr , can be determined as a function of to, y,
and the time, t - to' measured from the time of interruption.
(5-64)

The stress and strain patterns for this type of test are sketched in
Figure 5-24.
Meissner measured the ultimate recoil, Yoo' following steady simple shear of magnitude Yo for various shearing times for LDPE
"Melt I" [43]. He also measured the stress growth functions, (T+
and N7, and all three functions are shown in Figure 5-25. The
strain (y = yt) is used in place of the time as the independent
variable. We note that the maxima in the three curves occur at
different values of the strain (and of the time). At small strains, the
ultimate recoil is equal to twice the stress ratio, (N1/(T). It decreases with increasing strain reaching a minimum of (0.5 N1/(T) at
a strain of about 25 and then increases as the strain increases
further.

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

80

~x 3
.,....

60

215

I1l

-1

0..

6
()

w 2

40

II:

enen
w

II:
I-

W
l-

en

:2

20

--1

:::>

10
SHEAR STRAIN, y =

100

yto

Figure 5-25. Ultimate recoil during start-up flow, shear stress growth function and first
normal stress growth function for LDPE "Melt I." The shear rate is 2 S~l and the
temperature is 150C. Adapted from Ref. 43. Copyright 1975 by Steinkopff Verlag. Reprinted
by permission.

Wagner and Laun [40] found that data for ultimate recoil during
the start-up of steady shear could be plotted in a temperatureinvariant manner by using the usual shift factor, aT' for the shear
rate and by making the shear strain the independent variable rather
than the shearing time. This implies that:
(5-65)
Maxwell [44] has used recoil after a period of constant y as the
basis for a "melt index elasticity test" designed to provide a rapid,
simple evaluation of the elasticity of a melt?
5.5.3 Recoverable Shear Following Steady Simple Shear

The recoverable shear during steady simple shear is a function only


of the shear rate, and we represent this quantity as Yoo( y). It can be
considered as either Yoo(lT), defined by Equation 5-57, or the limiting
2A commercial instrument designed by Maxwell expressly for this test is made by Custom
Scientific, whose address is given in Appendix E.

216

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

value of 'Yoo( 'Y, y), which appears in Equation 5-65, for long duration
of the preceding start-up flow.
Laun [45] used a single-exponential damping function (Equation
3-72) together with Wagner's equation (3-64) to calculate 'Yoo( y). He
noted that the stress decay coefficients predicted by these equations
(Equations 5-42 and 5-43) indicate that the relaxation strengths are:
G.

g;(y)

(1 +

(5-44)

n'YAY

He then assumed that these same relaxation strengths govern the


recoil process and used the linear viscoelastic expression for 'Yoo
with G; replaced by g/ y) and TJo replaced by TJ( y). The result is:

'Yoo( y)

,T, ( . ) .
't'1 'Y 'Y
=

.2
n'Y

2 (") + - (
.)
TJ 'Y

\2

TJ 'Y,

(1

TJ;Il;

+ n'Y. A;)3

(5-66)

This can be contrasted with the prediction of the rubberlike liquid


model, which is:
(3-48)

The predictions of Equations 5-66 and 3-48 are compared with


Laun's data for LDPE "Melt I" [45] in Figure 5-26. Also shown (the
lowest curve) is the quantity ['l'1( y)y /2TJ( y)], i.e., the first term of
Equation 5-66. For low shear rates where the viscosity is independent of shear rate, the rubberlike liquid model is adequate, but at
higher shear rates it strongly overpredicts the recoverable shear. At
higher shear rates Equation 5-66 gives a good prediction, while one
half of the stress ratio, evaluated at the appropriate shear rate,
underpredicts the data. Thus, half the stress ratio is not a quantitative measure of recoverable shear in steady simple shear flow.
Laun [45] has also proposed the following empirical equation that
relates the recoverable shear, 'Yoo( y), to the loss angle measured in a

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

10

217

-2

10-3~----~----~----~----~----~----~--~
10- 4
10 3
10- 2
10- 1
100
101
10 2
te 3
SHEAR RATE.

Y (5)

Figure 5-26. Ultimate recoil of LDPE "Melt I" after steady shear. Also shown are the
predictions of Equations 3-48 (rubberlike liquid, straight line) and 5-66. The dashed curve is
the first term of Equation 5-66. Adapted from Ref. 45. Copyright 1986 by The Society of
Rheology. Reprinted by permission of John Wiley & Sons, Inc.

small amplitude oscillatory shear experiment:

I'oo{y)

cotan ~{w)[secant ~{w)]3

with w

(5-67)

Laun found that this expression was valid for a number of commercial resins, although he shows data only for high density polyethylene.
5.6 SUPERPOSED DEFORMATIONS

Large, high shear rate deformations alter the physical structure of a


melt by the processes of disentanglement and molecular orienta-

218

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

tion, and it is desirable to be able to monitor this reversible change


of state during a deformation. One technique that has been used to
do this is to measure the response of a small "probe" deformation
superposed on the larger one of interest. It is then assumed that the
small deformation has negligible effect on the change of state
resulting from the main flow. However, since such a change of
structure always results in deviations from linear viscoelasticity, the
Boltzmann superposition principle does not apply, and some coupling between the main and superposed flows is likely. Therefore,
the validity of a given superposed flow technique must be demonstrated by measuring the response to the main flow with and
without the superposed flow.
Another problem that arises is that the transducer used to
determine the stress must have a large enough range to accommodate the response to the main flow, but at the same time it must be
sensitive enough to measure the much smaller response to the
superposed flow.
5.6.1 Superposed Steady and OSCillatory Shear

Two modes of superposed steady and oscillatory shear have been


used. In the orthogonal mode, the oscillatory deformation is in the
X3 direction while the steady shear is in the Xl direction. The strains
in the two directions are as follows:
'Y21 =

mt

'Yo

'Y23

(5-68a)

In the parallel mode, both flows are in the

dV l
yet) == dx
2

(5-68b)

sin(wt)
Xl

direction, and:

Ym + 'YOW cos(wt)

(5-69)

Material functions for presenting the results of such tests have been

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

219

defined [46], and they have been used to study polymer solutions
[47-49].
5.6.2 Step Strain with Superposed Deformations

Isono and Ferry [50] used intermittent small amplitude oscillatory


shear to probe the relaxation of polyisobutylene following a larger
step strain.
McKenna and Zapas [51] used a small step strain, superposed
after a delay time, to probe the relaxation of a polyisobutylene after
a larger step strain. They concluded that it was not possible to
interpret the results of such a test without reference to a specific,
nonlinear viscoelastic model.
5.7 LARGE AMPLITUDE OSCILLATORY SHEAR

In Section 2.8, the use of small amplitude oscillatory shear to study


the linear viscoelastic behavior of liquids was described. The strain
is sinusoidal in time:

y{t)

Yo sin{wt)

(2-45)

and if the strain amplitude, Yo, is sufficiently small, the shear stress
is also sinusoidal:

a{t)

a o sin{wt + D)

(2-47)

The amplitude ratio (ao/yo) and the loss angle (0) are functions of
frequency but not of Yo.
As the amplitude is increased, the behavior will no longer be
governed by the Boltzmann superposition principle. One manifestation of nonlinearity is that Nt(t) becomes significant. In addition,
the stress is no longer sinusoidal, in which case there are higher
harmonics, and a(t) can no longer be described in terms of two
functions of frequency (Gd and 15 or G' and Gil).
Nevertheless, a strain amplitude-dependent complex modulus or
complex viscosity is sometimes used to describe nonlinear results in
the region of small deviations from linearity. For example, Ohta

220

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

et al. [52] used this method to represent the results of their study of
the microdomain structure of block copolymers. However, they did
not verify that the stress was sinusoidal. The algorithms usually
employed to calculate G' and G" from the u(t) signal generated by
a rheometer cannot discriminate between linear (sinusoidal) and
nonlinear responses and will produce values for these functions
even if the stress is not sinusoidal. For example, cross-correlation
gives only the first harmonic of the stress signal, even if higher
harmonics are present. It is essential to look at the unprocessed
stress signal to detect nonlinearity.
The fact that the shear stress is not sinusoidal complicates the
interpretation of data from large-amplitude tests. On the other
hand oscillatory shear has some advantages over flows such as step
strain, start-up and creep, in that a sudden change of the displacement, velocity or load is not required. Furthermore, because it is a
frequency domain rather than a time domain test, the amplitude of
the deformation ('Yo) and the time scale (l/w) can be independently varied.
It is of interest to construct a map showing the various regimes of
behavior that can be exhibited at various combinations of amplitude
and frequency. Pipkin [53] has suggested that a useful basis for such
a map is a plot of strain rate amplitude (w'Yo) versus frequency. A
diagram of this type is shown in Figure 5-27.
At very low frequencies, the deformation becomes indistinguishable from steady simple shear, because the shear rate variation with
time is very slow. Thus, in a zone along the left side of the diagram,
the behavior will be governed by the viscometric functions. If the
shear rate amplitude is also small, Newtonian behavior will be
exhibited, as shown in the lower left hand corner. The vertical line
that defines the region of viscometric flow is schematic only. It is
likely that it will bend to the right toward the top of the diagram,
reflecting the decrease in viscosity and effective relaxation time as
the shear rate amplitude increases.
In Pipkin's original diagram [53] Ayo is shown as a function of
Aw, where A is the relaxation time of the material. In this way, the
diagram is in terms of dimensionless variables. These dimensionless
groups have the following significance.
Aw

Deborah number

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

(;
.s;

.~

u..i

Nonlinear

Viscoelasticity

l-

::J

a.
~

l-

a:
z

<:

a:
Ien

i:

221

<'0

.s::
Q)

a:l

u:

<'0

iIi

.~

Qi
E
0

u
(/l

:>

- - . - Newtonian Behavior

Viscoelasticity

FREQUENCY, W

Figure 5-27. Pipkin diagram (Yo versus w) showing regimes of behavior for oscillatory shear.
(Not to scale; viscometric flow and linear viscoelasticity zones are actually quite narrow.)

This group is a measure of the extent to which elastic or memory


effects will playa role in the response of the fluid.
kyo = Weissenberg number
This group is a measure of the extent to which anisotropy, i.e.,
nonlinearity, will be exhibited in the response. Large amplitude
oscillatory shear is one of the few simple experiments in which
these two groups can be varied independently. Another strain
history that has this property is the up-down ramp in shear rate that
is used to produce "thixotropic loops" [54]. However, because there
is no uniquely important relaxation time for a polydisperse or
branched polymer, it is not possible to make quantitative use of
these dimensionless variables except for linear monodisperse
polymers.
As the frequency is increased, the stress will begin to lag behind
the strain, and the material will exhibit viscoelasticity. It is of

222

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

interest to locate the region in which this response will be linear.


Astarita and Jongschaap [55] have pointed out that basic concepts
of continuum mechanics can be used to show that at low frequencies it is the strain rate amplitude that should govern departures
from linearity, while at higher frequencies it is the strain amplitude
that governs the onset of nonlinear behavior. On the Pipkin diagram, the strain amplitude is constant along straight lines through
the origin, and one such line is shown in Figure 5-27 as the limit of
linear viscoelastic behavior at higher frequencies.
In the region just to the right of the line dividing the zone of
viscometric flow from that of nonlinear viscoelasticity, some authors
have reported their data in terms of storage and loss moduli that
depend on the strain amplitude [56] or the strain rate amplitude
[57] as well as the frequency. This implies that the stress signal has
no higher harmonics in this region, even though the behavior is
clearly nonlinear. However, there is no theory that predicts such a
behavioral regime, and it seems likely that the analog instrumentation used to acquire data in these experiments was unable to detect
the third and fifth harmonics as they began to appear. Philippoff
[58] was able to detect the third harmonic, and studied the onset of
nonlinearity in several polymer solutions.
As the stress amplitude increases, generally moving up the diagram, there is an increased likelihood of melt fracture, and this
phenomenon has been studied by Vinogradov et al. [59].
As frequency is increased, the behavior becomes more and more
elastic, and there is a zone along the right hand side of the diagram
in which the response is nondissipative. Coming back out of this
zone into the nonlinear viscoelasticity zone, it has been suggested
that there may be a region in which the stress is sinusoidal but the
amplitude ratio and loss angle depend on the strain amplitude.
Turning now to the problem of interpretation of data, the most
straightforward approach is to use a Fourier series representation
of the stress. For an isotropic material with fading memory, it can
be shown that the stress can be represented as follows:

u(t)

Yo

[G~(w, Yo)sin(nwt)

+ G;:(w, Yo)cos(nwt)] (5-70)

n=l
n odd

According to a number of viscoelastic equations of state, the


nonlinear storage and loss moduli can be expanded in the odd

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

223

powers of the strain amplitude, and the stress can thus be represented as follows:
00

a(t) = Yo

L L

[G~p(w)sin(nwt)

+ G';p(w)cos(nwt)] (5-71)

p=l n=l
p odd n odd

Algorithms for calculating the discrete Fourier transform (FFT


programs) make it easy to extract the Fourier components from a
stress signal if noise is not a problem [60]. Onogi et al. [61,62] have
analyzed the behavior of carbon black-filled polystyrene solutions in
terms of Fourier coefficients and have studied the temperature
dependence of these material constants.
The use of Fourier analysis provides a complete mathematical
description of the response to large amplitude oscillatory shear.
However, it yields a rather complex representation of the behavior
in terms of plots of G~(w, Yo) and G~'(w, Yo) for as many harmonics
as can be identified in the response. A representation of the data
that is more amenable to rapid qualitative evaluation is the use of a
closed loop plot of stress versus strain or stress versus rate of strain.
For the case of linear viscoelastic behavior these loops are ellipses,
but when higher harmonics are present, the ellipse is distorted.
Tee and Dealy [63] found that for molten thermoplastics at
modest frequencies, the stress versus rate-of-strain loops were more
distinctive than the stress-strain loops. Figures 5-28 and 5-29 show
a - y loops for a low density polyethylene and a polystyrene,
respectively, at several frequencies and amplitudes [63].
One interesting feature of the stress response is the energy
dissipation per cycle per unit volume. This is given by:

P=

f ady = j 2Tr/Wa(t)y(t) dt
o

(5-72)

This is the area inside the stress-strain loop. Using the Fourier
series for a(t) (Equation 5-70) it can be further shown [64] that the
dissipation is associated with the first harmonic:

P(wYo) = 7Ty~GO(W,yo)

= 7T

L
p=l
odd

G'{p(whg+ 1

(5-73)

224

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

;r

V
V V:

. . .v

./.

/
.....

/
...........

/'

/:

...........

Figure 5-28. Stress versus shear rate loops for oscillatory shear of low density polyethylene.
Left: ill = .94 s-t,"Yo = 2.2,0"0 = 1.3 X 10 4 Pa. Right: ill = .63 s-t,"Yo = 9.6,0"0 = 2.1 X 10 4
Pa. Adapted from Ref. 63. Copyright 1975 by The Society of Rheology. Reprinted by
permission of John Wiley & Sons, Inc.

While the normal stress differences are zero for linear viscoelastic
behavior, they are nonzero for large amplitude oscillatory shear.
Lodge [65] has shown that the rubberlike liquid model predicts that
the first normal stress difference is given by:

N 1{t,w,'Yo)

= 'Y~[G"{w)

- B{w)cos{2wt) - C{w)sin{2wt)]
(5-74)

;.........-

"7

L-

....-- V

--- /.. L

7'
........

l? f--

---

./:

I--- ;7

V
y

Figure 5-29. Stress versus shear rate loops for oscillatory shear of a molten polystyrene. Left:
= 4.4 s-t, 'Yo = 5,0"0 = 7.1 X 10 4 Pa. Right: ill = 2.5 s-', 'Yo = 10, 0"0 = 6.1 X 10 4 Pa.
Adapted From Ref. 63. Copyright 1975 by The Society of Rheology. Reprinted by
permission of John Wiley & Sons, Inc.

ill

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

225

This suggests that when nonzero values of Nl are first detected (as
Yo is increased) Nl will be sinusoidal with an amplitude proportional to Y5, a frequency that is twice the strain frequency, and a
positive average value.
Predictions of (T(t, w, (To) for several other constitutive equations
have been published [66-71].
5.8 EXPONENTIAL SHEAR: A STRONG FLOW

Exponential shear flow is defined as follows

yet)

A(e at

1)

(5-75)

where a is the exponential rate constant and A is the strain scale


factor. This deformation is of special interest, because it tends to
generate a high degree of molecular stretching and is thus a
"strong" flow [72]. All the other shear flows described in this
chapter are "weak" flows, in that there is only a rather mild
tendency to stretch out molecules.
The strain rate is given by:

yet)

=aAe at

(5-76)

The response will always become nonlinear at some value of t since


the shear rate increases exponentially. However, at the beginning of
an experiment we expect the response to obey the Boltzmann
superposition principle, which predicts that:

(5-77)
In the limit of a

0, we have:
(5-78)

226

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

For a Maxwell fluid with relaxation time A, we have:


(T(t) =

aA7J

eat

(1 + aA)

[1 _ e-(l+aA)t/A]

(5-79)

Thus, the stress increases without limit. It is not possible to define a


material function for this flow that becomes independent of a for
small values of a, because unlike steady simple shear this is not a
flow with constant stretch history.
However, it is possible to define a material function that has a
finite limiting value at large t for the case of linear behavior [69].
This is the exponential viscosity, 7J e (t, a, A), defined as follows:
7Je(t, a, A) ==

(T(t,a,A)
A at
a e

(5-80)

For the Maxwell fluid, this material function is independent of A


and behaves as follows:
(5-81)
For large values of

t/

A the limiting value is:


(5-82)

Thus, the use of this material function facilitates the comparison of


experimental data with linear behavior. While the entire linear
curve cannot be obtained experimentally, a hypothetical curve can
be computed by use of Equation 5-81, generalized to accommodate
a discrete spectrum.
We note that the denominator in (5-80) is simply the shear rate,
y. It has been suggested [72] that in the experimental determination
of the exponential viscosity, 7Je should be calculated by dividing the
measured shear stress by the measured shear rate. This eliminates

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

227

40~------------------------------------------~

30

10

o~------~------~--------~------~------~
4
16
8
20
o
12
TIME (5)

Figure 5-30. Exponential viscosity of LDPE "IUPAC X" Oowest curve) compared with
theoretical linear response (top curve) and the prediction of Wagner's equation. Adapted
from Ref. 75. Copyright 1989 by The Society of Rheology. Reprinted by permission of
John Wiley & Sons, Inc.

cumulative error due to deviations of y(t) from (5-75) at the start of


the experiment.
While rotational rheometers are not useful for the generation of
exponential shear, sliding plate rheometers have been found to be
well suited to this type of test [73,74]. Figure 5-30 shows experimental data for a low density polyethylene [75] together with the
theoretical curve for the linear viscoelastic response. Also shown is
the prediction of Wagner's equation, using a damping function
obtained by shifting step shear strain data. We note that the data
start out on the linear curve and then fall below it. At large times,
the curve decreases sharply, indicating that this flow generates very
rapid disentanglement at these times. The most interesting part of
the curve from the point of view of nonlinear viscoelasticity is the
region surrounding the maximum.
Doshi and Dealy [72] have proposed an alternative definition for
a material function, which emphasizes the role of molecular stretching and orientation. It is based on the use of the principal compo-

228

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

nents of the stress and strain rate and is defined as follows:

(5-83)

The rubberlike liquid model predicts that this function rises above
the theoretical curve for linear behavior at shorter times as a is
increased [72].
5.9 USEFULNESS OF TRANSIENT SHEAR TESTS

In the introduction to this chapter, we discussed the reasons for


carrying out transient shear tests of melts. Here we summarize the
uses of transient shear data.
In polymer engineering, nonlinear viscoelastic properties may be
useful in two ways. First they are more sensitive to certain features
of the molecular structure than linear and steady shear properties.
In addition, a nonlinear test can sometimes be a useful indicator of
the way a melt will perform in a particular processing operation.
In polymer science, the use of large transient deformations
makes it possible to learn about the strengths and weaknesses of
constitutive equations and molecular models. Finally, observations
of unexpected phenomena can inspire the formulation of new
continuum and molecular theories of rheological behavior.
REFERENCES

1.
2.
3.
4.
5.
6.
7.
8.
9.
10.

J. Meissner, Pure and Appl. Chem. 42:553 (1975).


H. H. Winter, Pure and Appl. Chem. 55:943 (1983).
J. L. White and H. Yamane, Pure and Appl. Chem. 59:193 (1987).
C. M. Vrentas and W. W. Graessley, J. Non-Newt. FI. Mech. 9:339 (1981).
L. J. Zapas, J. Res. N. B. S. 75a:33 (1971).
H. M. Laun, Rheol. Acta 17:1 (1978).
K. Osaki, K. Nishizawa and M. Kurata, Macromolecules 15:1068 (1982).
Y. K. Osaki, E. Takatori and M. Kurata, Macromolecules 20:1681 (1987).
K. Osaki and M. Kurata, Macromolecules 13:671 (1980).
M. Fukada, K. Osaki and M. Kurata, J. Polymer Sci. Polym. Phys. 13:1563
(1975).

TRANSIENT SHEAR FLOWS USED TO STUDY NONLINEAR VISCOELASTICITY

229

11. C. M. Vrentas and W. W. Graessley, J. Rheo!. 26:59 (1982).


12. Y. Einaga, K Osaki, M. Kurata, S. Kimura, N. Yamada and M. Tamura,
Polymer J. 5:91 (1973).
13. A. V. Tobolsky and K Murakami, J. Polymer. Sci. 40:443 (1959).
14. K Osaki, S. Ohta, M. Fukada and M. Kurata, J. Polymer Sci. Polym. Phys.
14:1701 (1976).
15. x.-L. Luo and R. I. Tanner, J. Non-Newt. Fl. Mech. 22:61 (1986).
16. K Osaki, N. Bessho, T. Kojimoto and M. Kurata, J. Rheo!. 23:617 (1979).
17. K Osaki, S. Kimura and M. Kurata, J. Polymer Sci. 19:517 (1981).
18. H. M. Laun, J. Rheol 30:459 (1986).
19. S. Kimura, K Osaki and M. Kurata, J. Polymer Sci. 19:151 (1981).
20. K Osaki, S. Kimura and M. Kurata, J. Rheol. 25:549 (1981).
21. K Osaki and M. Kurata, J. Polymer Sci. Polym. Phys. 20:623 (1982).
22. M. H. Wagner and S. E. Stephenson, Rheol Acta 18:463, (1979).
23. I. Broker, Rheol. Acta 25:501, (1986).
24. W. W. Graessley, W. S. Park and R. L. Crawley, Rheo!. Acta 16:291 (1977).
25. M. H. Wagner, Rheol. Acta 15:136 (1976).
26. J. Meissner, J. Appl. Polymer Sci. 16:2877 (1972).
27. E. V. Menezes and W. W. Graessley, J. Polymer Sci. Polym. Phys. 20:1817
(1982).
28. J. Meissner, Ann. Rev. Fluid Mech., 17:45 (1985).
29. R. A. Stratton and A. F. Butcher, J. Polymer Sci. Polym. Phys. 11:1747 (1973).
30. H. A. Pohl and C. G. Gogos, J. App!. Polymer Sci. 5:67 (1961).
31. J. M. Dealy and W. K W. Tsang, J. App!. Polymer Sci. 26:1149 (1981).
32. W. K W. Tsang and J. M. Dealy, J. Non-Newt. F!. Mech. 9:203 (1981).
33. D. E. Hanson, Polym. Eng. Sci. 9:405 (1969).
34. B. Maxwell, E. J. Dormier, F. P. Smith and P. P. Tong, Polym. Eng. Sci.
22:280 (1982).
35. A. Rudin and H. P. Schreiber, Polym. Eng. Sci. 23:422 (1983).
36. G. Ritzau, Int. Polym. Proc. 1:188 (1987).
37. P. K Agarwal and D. J. Plazek, J. Appl. Polym. Sci. 21:3251 (1977).
38. H. M. Laun and J. Meissner, Rheo!. Acta 19:60 (1980).
39. K Nakamura, C. P. Wong and G. C. Berry, J. Polym. Sci. Polym. Phys.
22:1119 (1984).
40. M. H. Wagner and H. M. Laun, Rheo!. Acta 17:138 (1978).
41. H. M. Laun, Prog. Colloid Polym. Sci. 75: (1987).
42. J. Meissner, Rheo!. Acta 14:471 (1975).
43. J. Meissner, Rheo!. Acta 14:201 (1975).
44. B. Maxwell, SPE Tech. Papers 33:1659 (1989).
45. H. M. Laun, J. Rheo!. 30:459 (1986).
46. J. M. Dealy, J. Rheo!. 28:181 (1984).
47. H. c. Booij, Rheol. Acta 7:202 (1968).
48. R. I. Tanner and G. Williams, Rheo!. Acta 10:528 (1971).
49. K Osaki, M. Tamura, M. Kurata and T. Kotaka, J. Phys. Chem. 69:4183
(1965).
50. Y. Isono and J. D. Ferry, J. Rheo!. 29:273 (1985).

230

51.
52.
53.
54.
55.
56.
57.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

G. B. McKenna and L. J. Zapas, Polym. Eng. Sci. 26:725 (1986).


Y. Ohta, T. Kojima, T. Takigawa and T. Masuda, J. Rheol. 31:711 (1987).
A. C. Pipkin, Lectures in Viscoelastic Theory, Springer-Verlag, N.Y., 1972.
J. Greener and R. W. Connelly, J. Rheol. 30:285 (1989).
G. Astarita and R. J. J. Jongschaap, J. Non-Newt. Fl. Mech. 3:281 (1977-78).
I. F. Macdonald, B. D. Marsh and E. Ashare, Chern. Eng. Sci. 24:1615 (1965).
G. V. Vinogradov, Yu. Yanovsky and A. I. Isayev, J. Polymer Sci. A-2 8:1239

(1970).
58. W. Philippoff, Trans. Soc. Rheol. 10:317 (1966).
59. G. V. Vinogradov, A. I. Isayev and E. V. Katsyutsevich, J. Appl. Polym. Sci.
22:727 (1978).
60. H. Ramirez, The FFT-Fundamentals and Concepts, Prentice Hall, Englewood
Cliffs, N.J. 1985.
61. S. Onogi, T. Masuda and T. Matsumoto, Trans. Soc. Rheol. 14:275 (1970).
62. T. Masumoto, Y. Segawa, Y. Warashina and S. Onogi, Trans. Soc. Rheol.
47:(1978).
63. T. T. Tee and J. M. Dealy, Trans. Soc, Rheol. 19:595 (1975).
64. S. Onogi and T. Matsumoto, Polymer Eng. Rev. 1:45 (1981).
65. A. S. Lodge, Elastic Liquids, Academic Press, NY (1964).
66. R. I. Tanner, Trans. Soc. Rheol. 12:155 (1968).
67. H.-C. Yen and L. V. McIntyre, Trans. Soc. Rheol. 16:711 (1972).
68. D. S. Pearson and W. E. Rochefort, J. Polym. Sci. Polymer Phys. 20:83 (1982).
69. E. Helfand and D. S. Pearson, J. Polym. Sci. Polym. Phys. 20:1249 (1982).
70. X. J. Fan and R. B. Bird, J. Non-Newt. Fl. Mech. 15:341 (1984).
71. W. K. W. Tsang and J. M. Dealy, J. Non-Newt. Fl. Mech. 9:203 (1981).
72. S. R. Doshi and J. M. Dealy, J. Rheol. 31:563 (1987).
73. B. Zlille, J. J. Linster, J. Meissner and H. P. Hlirlimann, J. Rheol. 31:583
(1987).
74. N. Sivashinsky, A. T. Tsai, T. J. Moon and D. S. Soong, J. Rheol. 28:287
(1984).
75. T. Samurkas, R. G. Larson and J. M. Dealy, J. Rheol. 33:539 (1989).

Chapter 6
Extensional Flow Properties
and Their Measurement
6.1 INTRODUCTION

For deformations that are either very small or very slow, the theory
of linear viscoelasticity is a unifying concept that provides relationships between the material functions that are determined using
various types of deformations. For example, this theory tells us that
in start-up flow, the shear stress growth coefficient, 11 +(t), is independent of shear rate. Furthermore, it provides a simple relationship between this material function and the tensile stress growth
function, 11;(t), that is measured at the start-up of steady simple
extension:
(6-1)

Thus, as long as the total strain or the maximum strain rate that
occurs during a particular deformation is very small, no new information is obtained from the use of an extensional flow, once the
linear viscoelastic behavior has been established by use of a shearing deformation. A corollary of this statement is that the response
of a melt to any small or slow extensional flow can be calculated
from a material function determined using a small or slow shearing
experiment.
For large, rapid deformations, however, there is no unifying
theory, and the response of a melt depends on the size, rate and
kinematics of the deformation. In Chapters 4 and 5, the effects of
strain rate and strain magnitude on the responses to various shearing deformations were examined. In the present chapter we will see
231

232

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

EJ

Figure 6-1. Separation of two particles located on the same streamline during simple
extension.

how a variation in the kinematics of the deformation, i.e., going


from shear to extension, influences the response. However, even at
large strain rates, the initial portion of the curve of 'TlI(t, i) should
follow Equation 6-1, since the total strain is still small at small
values of t.

6.2 EXTENSIONAL FLOWS

The simplest definition of an extensional flow is that it is a deformation that involves stretching along streamlines. For example, in
uniaxial (simple) extension, if the Xl axis is aligned with the principal stretching direction, we note that two fluid particles located on
this axis will get further apart as the deformation proceeds, as
shown in Figure 6-1. By contrast, we note that in simple shear, two
particles located on any line parallel to a velocity vector, i.e., on a
given streamline, do not get further apart as the deformation
proceeds, as shown in Figure 6-2.
This is not to say that no stretching occurs in a simple shear
deformation. In fact, all deformations involve stretching. However,
in shear flows, one must consider material particles located on
different streamlines to see the stretching, as shown in Figure 6-3.

EJ

Figure 6-2. Displacements of two particles located on the same streamline during simple
shear. These are equal; there is no stretching along the line joining these two particles.

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

233

Figure 6-3. Separation of two particles located on different streamlines during simple shear.
There is stretching along the line joining these particles.

For simple shear flow, it is common practice to use a coordinate


system in which the Xl axis is parallel to the velocity vector, and
there is no motion in the X3 direction. For this choice, the sLain
rate components are:
'}'

o
o

~]

(6-2)

We note that for this choice of coordinate system there are no


diagonal components, although if the coordinate axes are rotated,
diagonal components can appear. For the simplest extensional
flows, often called "shear free" flows, it is convenient to use a
coordinate system in which the axes are oriented in the directions
of the principal strain axes. For this choice, the strain rate components always have the following general form:

o
(6-3)
where at, a 2 , and a 3 have units of reciprocal time, and at + a 2

a 3 = 0 for an incompressible material. We note that there are no

off-diagonal components. Sometimes the form of the matrix containing the components of the rate of strain tensor is used as a basis
for distinguishing between shear and extensional flows. However,
the form of these matrices depends on the choice of coordinate
system, so this is not a rigorous classification scheme. However, it is
true that for extensional flows the principal axes for stress and
strain are the same, whereas this is not the case for shear flows

234

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

t> ~---------------tJ
Figure 6-4. Effect of simple extension on an initially cubical fluid element.

(except in the case of Newtonian fluids). In any event, the most


obvious characteristic of an extensional flow is that there is stretching along streamlines.
The simplest extensional flows, with regard to describing both the
deformation and the resulting stresses, are uniform, shearfree flows,
in which the strain rate is the same for every material element, and
there is no relative rotation of perpendicular axes fixed in a fluid
element. It is convenient to use the principal strain directions as the
coordinate axes, and in this case the rate of strain tensor has only
diagonal components, as shown in Equation 6-3. The shear free
flow that has been most used in experimental rheology is uniaxial
extension, also called simple extension, which is an axisymmetric
flow with stretching in the direction of the axis of symmetry. The
effect of uniaxial extension on the shape of an initially cubic fluid
element is sketched in Figure 6-4.
There is another axisymmetric extensional flow, and this is biaxial
extension. In this deformation there is compression along the axis
of symmetry and stretching in the radial direction, as shown in
Figure 6-5. This exhausts the possibilities for axisymmetric deformations, but there are many possible non-axisymmetric shear free
flows. The simplest of these is "planar" extension, in which the
velocity in one Cartesian direction is zero, as shown in Figure 6-6.

Figure 6-5. Change in shape of a fluid element during biaxial extension.

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

235

Figure 6-6. Change in shape of a fluid element during planar extension.

This type of flow can be approximated, for example, by stretching a


thin-walled tube in its axial direction while supplying sufficient
internal pressure to keep the diameter constant (see Section 6.6).
It is of interest to have some understanding of the changes in
molecular conformation and organization generated by different
types of deformation. For uniaxial extension, the tendency is to
align molecules with the axis of symmetry, i.e., in a single direction.
By contrast, in biaxial extension the tendency is to align molcecules
in the radial direction, but this implies a symmetrical distribution of
alignments in planes perpendicular to the axis of symmetry. In
other words, within one of these planes, there is no overall preferred orientation in anyone direction. Planar extension is between
these two extremes, with a tendency to orient molecules parallel to
the X I -X 3 plane, and to some degree in the Xl direction.
A related issue is the question of "strain hardening" or "extension thickening." This concept is based on a comparison of the
stress in start-up flow with that predicted by the theory of linear
viscoelasticity. If the stress increases more rapidly than would be
predicted by the linear theory, the material is said to be "strain
hardening" suggesting an analogy with the strain hardening of
metals. However, the latter phenomenon is associated with plastic
flow or yielding rather than viscoelasticity, and a better term for
melts is "extension thickening."
In Figure 6-7 this type of behavior is illustrated by Curve 2,
where Curve 1 is the linear viscoelastic response. By contrast, strain
softening or extension thinning, refers to the behavior shown by
Curve 3, where the normalized stress, i.e., the tensile stress growth
coefficient, 7];, drops below the linear viscoelastic curve at some
value of t before which the strain is sufficiently small that linear
behavior is exhibited. In general, polymers with a significant degree

236

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

________________________________

TIME

Figure 6-7. Types of behavior exhibited by polymeric liquids during tensile start-up flow.
Curve 1 shows the response predicted by the linear theory of viscoelasticity. Curve 2 shows
extension thickening (strain hardening) behavior, while curve 3 shows extension thinning
behavior.

of long chain branching exhibit marked extension thickening, while


linear materials usually exhibit extension thinning. An alternative
way of classifying extensional flow phenomena is to compare a
material's response to steady strain rate stretching with the prediction of the Lodge rubberlike liquid equation presented in Chapter
3. Looked at in this way, all melts exhibit extension thinning.
Nearly all melt processing operations involve some kind of extensional flow, as there is always stretching along streamlines at one or
more stages of the process. For example, whenever there is a
converging flow, as in a die or nozzle, there is acceleration, and this
implies the presence of uniaxial or planar extension, depending on
the flow geometry. In a diverging channel, there will be extension in
directions normal to the main flow direction. However, a convergence or divergence is generally the result of a change in the cross
section of a flow channel, and if the melt adheres to the wall of the
channel, near the wall the kinematics will be primarily shear rather
than extension.
Melt spinning is an example of a uniaxial extensional flow in
which the strain rate changes along the flow path. Parison sag in the
blow molding process is another example of uniaxial extension,

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

237

while parison inflation is intermediate between planar extension


and biaxial extension. The blown film process also involves flow in
this category, while in film casting the deformation is intermediate
between uniaxial extension and planar extension. In film extrusion
operations, the "drawability" of the melt and the stability of the
flow in the melt region downstream of the die are of central
importance. It is thought that these two phenomena are related to
extensional flow properties. Drawability, the ability of a melt to be
stretched into a thin film without breaking, is associated with
extension thinning (strain softening) behavior, while good bubble
stability is associated with extension thickening (strain hardening)
behavior.
6.3 SIMPLE EXTENSION

Simple (uniaxial) extension is a uniform deformation in which the


velocity distribution is given by:

(6-4a)
(6-4b)
(6-4c)
Since the flow is axisymmetric it can also be conveniently described
using cylindrical coordinates:

(6-5a)
Vr =

-1/2er

(6-5b)

The Hencky strain rate, e, first introduced in Equation 1-7, is


uniform in space but can be a function of time. The components of
the rate of deformation tensor are given by Equation 1-48, shown
below, and the components of the Cauchy and Finger tensors are
given by Equations 3-3 and 3-4.

o
-I::

(1-48)

238

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

In contrast to the situation in simple shear, the principal stresses


for extensional flow are in the same directions as those for the
strain rate. Thus, the stress tensor also has only diagonal components:

(6-6)
For incompressible fluids, only the differences between normal
stress components have rheological significance, and the "net tensile stress" in simple extension is defined as:

(6-7)
Thus, in a tensile flow experiment, one measures the tensile stress
(TE(t) for a prescribed strain history E(t). Alternatively the strain

resulting from a prescribed stress can be measured.


When the Hencky strain rate, 8, is constant (or if it increases with
time) simple extension is a "strong flow" in the sense that it tends
to generate a high degree of molecular stretching [1]. In addition,
Larson [2] suggests that it is a strongly aligning deformation. Thus,
by comparison with steady simple shear, a melt's response to simple
extension should reveal clearly the role of kinematics in nonlinear
viscoelastic behavior.
Indeed, a great deal has been learned about this question from
the study of the response of melts to simple extension. However,
the maximum strain rates that can be reached in the extensional
rheometers used to date are quite low, in fact far below those that
occur in melt processing operations. Furthermore, a high degree of
skill and care is required to obtain reliable extensional flow data,
and extensional rheometers have not come into general use for
resin evaluation.
6.3.1 Material Functions for Simple Extension

Step strain in extension can, in principle, be used to determine the


nonlinear tensile relaxation modulus, E(l, E), defined in Equation
2-2. This can then be used to determine the damping function, h(E).

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

239

From Equation 3-64, it can be shown that:

E(t, s)

e 2e _ e-e

---h(s)G(t)

(6-8)

However, this test is difficult to perform, because fracture or


necking nearly always occur. The tests that have been found most
useful are start-up of steady simple extension, tensile creep, and
unconstrained recoil following simple extension.
In start-up flow, the sample, initially in its rest state, is subjected
to steady simple extension at a rate i starting at time t = O. The
tensile stress growth coefficient is defined as follows:

(6-9)
At longer times, it is generally assumed that the stress will approach
a limiting constant value, which can be used to calculate the tensile
viscosity, TlE(i):

(6-10)
For the special case of linear viscoelastic behavior the tensile
stress growth coefficient reduces to a function of time alone and
becomes equal to three times the shear stress growth coefficient,
TI +(t), as shown by Equation 6-1. Thus, when start-up data are
being plotted in the form of curves of TIt versus t with i as a
parameter, the linear behavior should be shown for comparison on
the same graph. If the extensional rheometer used is not capable of
operation at a strain rate sufficiently low that linear behavior is
observed, the linear curve can be calculated by use of Equation 2-92
as follows:
limTlt(t,i) = Tlt(t) = 3T1+(t) = 3j tG(s)ds

i~O

(6-11)

The comparison with linear behavior has several advantages.


First it serves as a partial test of the validity of the nonlinear
results, since at short times or low strain rates the data should

240

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

follow the linear curve. For this reason, a plot of 7]t(t, i) is often
shown on the same graph with 37] +(0. Furthermore, the comparison with linear behavior provides a basis for describing the nonlinear behavior, i.e., the extent of the deviation from linear behavior is
a useful way of summarizing results without reference to a specific
nonlinear theory.
If a more quantitative description of the nonlinear behavior is
described, the damping function, h(e), can be determined. Wagner
[3] has derived the following equation for this purpose:

(6-12)

The argument of all the functions is the Hencky strain, e, and this
can be related to the usual time variables as follows:
t

t'

= eli

(6-13)

e'li

(6-14)

If a steady simple extension is suddenly halted after the stress has

reached a steady state level, the tensile stress decay coefficient can
be determined from the stress relaxation that ensues:
_
.
(TE(t)
7]E(t,e) == - . e

(6-15)

However, this test is little used because of the difficulty of achieving


a steady stress level while the sample is still homogeneous in cross
section.
Corresponding to the tensile creep compliance of linear viscoelasticity (Equations 2-43 and 2-44) is the nonlinear tensile creep
compliance, D(t, (TE):
(6-16)

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

241

At times sufficient for a constant strain rate to be achieved, D(t, (FE)


becomes linear with time and:
(6-17)

where Di(FE) is the steady state rensile compliance, and YJE is the
tensile viscosity value corresponding to the stress (FE' i.e., at a strain
rate given by:
.

(FE

=--

YJE(i)

(6-18)

At any time, to' during a start-up or creep experiment, the stress


can be eliminated and the recoil strain /t - to, to) measured as a
function of time (t - to) elapsed since the stress removal. The
ultimate tensile recoil or recoverable strain is 00 where
(6-19)

For a start-up experiment 00 will also be a function of i, while for a


creep experiment it will be a function of (FE'
6.3.2 Experimental Methods

Extensive reviews of experimental methods for determining extensional flow properties have been published [4-6]. The measurement
of extensional flow properties is much more difficult than the
measurement of shear flow properties for the following reasons:
1. There is no un deformed material surface that can be used to
support the sample or over which the deforming stress can be
applied. Since a solid surface cannot be used to support the
sample, it is necessary to suspend the sample in a bath of
neutrally buoyant oil or to float it on oil, unless the temperature is only slightly above the softening point.
2. The tensile force that can be transmitted from a wetted solid
surface to a melt is inadequate to supply the deforming stress.

242

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Various stratagems have been used to overcome this difficulty,


but the two methods now used in most cases are the rotary
clamp and the application of an adhesive.
3. The sample changes its shape during the experiment, and the
study of the response to large Hencky strains implies a very
large increase in length. For example, a Hencky strain of 5
corresponds to a stretch ratio of about 148, i.e., at a strain of
5, the sample length is 148 times its initial length! This means
that a long oil bath must be used if large strains are to be
reached. This makes more difficult the precise control of
temperature over the entire length of the sample.
4. For materials in which there is little or no extension thickening, any non uniformity in cross section of the sample will be
amplified during stretching, leading to nonuniform deformation and ductile failure. Such a situation makes it especially
crucial that the sample be perfectly homogeneous and that the
temperature be uniform along the sample length.
For these reasons extensional rheometers have proven to be of
limited use for the routine evaluation of molten plastics. Furthermore, many of the data that have been published are open to
question. In a few laboratories, extraordinary care has been taken
to ensure the reliability of the measurements, and the presentation
of melt data in a later section of this chapter will focus attention on
the contributions from these laboratories.
In the present section we will consider three basic types of device
designed to generate uniform simple extension; these are the extensional creepmeter, the "rotary clamp" rheometer of Meissner, and
the "universal extensional rheometer" of Miinstedt. At the end of
the section the usefulness of semi-empirical test methods involving
converging flow and extrudate drawing will be examined.
The simplest extensional rheometers are creepmeters in which
the tensile stress is applied by means of a weight suspended from a
specially designed cam mounted on a pulley. An early use of this
technique for molten thermoplastics was reported by Cogswell [7].
An improved design that has some advantages over other extensional creepmeters is that of Miinstedt [8], which is shown in Figure
6-8. The use of a vertical oil bath improves temperature uniformity,
although it prevents the floating of the sample on the oil and makes

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

243

AIR BEARING

CAM
DISPLACEMENT
TRANSDUCER

SAMPLE --Io/y--41

LOAD

-Ht-- OIL BATH

HEATING FLUID -

Figure 6-8. Extensional creepmeter designed by Miinstedt. Adapted from Ref. 8, Copyright
1975 by Steinkopff Verlag. Reprinted by permission.

necessary the matching of the densities of the melt and the oil. The
tensile stress is applied by means of an adhesive, and the elongation
is monitored electronically. This rheometer has been used to study
polystyrene [8] and LDPE [9].
In an extensional creepmeter, the stress, and thus the strain rate,
are limited on the low side by friction in the air bearing and on the
high side by the inertia of the moving components. This problem, as
well as the limitation on the maximum strain that is imposed by the
length of the oil bath, were avoided by Meissner [10, 11], who
developed an extensional rheometer based on the use of the
"rotary clamp." The principle of operation of this device is shown
in Figure 6-9. The two gear-like rotors turn at the same speed but
in different directions, thus generating a constant velocity in the

244

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 6-9. Extensional rheometer developed by Meissner. Zl and Zz are rotary clamps, M
is the drive motor for Zl' P is the sample, S is the leaf spring, and L is an LVDT to sense the
deflection of the spring. Arrows indicate cutters used for recoil measurements. Adapted from
Ref. 11. Copyright 1971 by Steinkopff Verlag. Reprinted by permission.

sample at a fixed point in space. From Equation 6-4a we see that


the Hencky strain rate is equal to the velocity of the melt passing
through the rotors divided by the distance from the rotors to a
vertical plane where the axial velocity is zero. The sample floats on
the surface of an oil bath. Remotely controlled scissors are used to
cut segments from the sample after stretching, to permit the measurement of the ultimate tensile recoil. Meissner used a second set
of slowly turning rotors at the other end of the sample to provide a
balancing tensile force, while Laun and Miinstedt [12] used an
adhesive to fasten this end of the sample to a fixed end-plate
submerged in the oil bath. They also made several other improvements to this apparatus and used it to study the stretching of LDPE
up to Hencky strains of 6. At the very large stretch ratio associated
with this strain, surface tension makes a significant contribution to
the tensile stress at low strain rates, and Laun and Miinstedt [12]
have presented an analysis of the effect of surface tension on
experimental results.
A number of extensiometers have been designed around the use
of a sample that is clamped at both ends, with one clamp fixed in
position and the second displaced by means of a servomotor [4, 6].
The configuration that has been most widely used is that proposed
by Miinstedt [13], the essential features of which are shown in
Figure 6-10. The sample is mounted in a vertical oil bath. The

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

245

ANGULAR DISPLACEMENT
TRANSDUCER
rNr-------SERVO MOTOR
WIND-UP DISC
-TAPE

__-THERMAL JACKET

-1--1---

.J,Io.--t--1f--_

OIL BATH
SAM P LE
FORCE TRANSDUCER

Figure 6-10. Extensional rheometer designed by Miinstedt. Adapted from Ref. 13, Copyright
1979 by John Wiley & Sons, Inc. Reprinted by permission.

bottom end of the sample is glued to an end plate coupled to a load


cell, while the top end plate is coupled to the end of a flexible band
that winds onto a drum driven by a DC servomotor. Only a very
small sample is required, and both constant strain and constant
stress experiments are possible. Commercial versions of this instrument have been offered by Rheometries and by G6ttfert. An
extensional rheometer similar in some respects to the Miinstedt
design but that can be used to make simultaneous birefringence
measurements has been developed by Muller [14] and is available
commercially from Metravib.
All of the extensional rheometers described above are limited to
use at strain rates less than about 1 s -1. This severely restricts their
use in the study of nonlinear viscoelasticity and in the simulation of
conditions typical of commercial melt forming operations. It is
possible to overcome this restriction if one gives up the use of a

246

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

uniform deformation in which the strain history of the sample is


precisely controlled. Specifically, the two flows that have been used
to generate uniaxial extension at higher strain rates are extrudate
drawing (melt spinning) and pressure-driven flow in a converging
channel.
In an extrudate drawing test, melt is forced through a capillary,
usually by means of a piston moving in a cylindrical reservoir. A
capillary rheometer is convenient for this purpose. The extruded
filament cools by exposure to ambient air, is drawn down by means
of a motor-driven drum, and the tensile force in the filament is
determined by measuring the vertical force on the rotating drum
(see Figure 6-11). Alternatively, the torque required to rotate the
drum can be measured. The test procedure usually used involves

I
_O(x)

Figure 6-11. Melt spinning apparatus used to determine melt strength.

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

247

increasing the wind-up speed and noting the force level when the
filament breaks [15]. This is called the "melt strength." This property has been found to be related to the performance of polyethylenes in the film blowing process [16, 17] and the sheet extrusion
process (see Section 17.5). Meissner [18] has shown how the rotary
clamp that he developed to generate homogeneous simple extension can be used to draw down the extrudate from a capillary and
thus measure the melt strength. The G6ttfert Rheotens is a commercial device based on this concept.
While the melt strength test shows promise as an empirical
measure of melt quality, it does not yield well-defined rheological
properties, because neither the strain nor the temperature are
uniform in the filament. The melt first experiences a shear deformation in the capillary and then moves into a zone in which the
deformation is primarily uniaxial extension, but the strain rate
varies continuously with distance from the capillary exit up to the
point where the melt becomes sufficiently cool that no further flow
occurs. Wissbrun [19] has pointed out that these factors make it
impossible to determine well-defined rheological material functions
using the melt strength test.
This flow has been analyzed in some detail by Bayer [20]. Using
some of Bayer's ideas, Laun and Schuch [21] derived equations for
calculating an apparent extensional viscosity at the take-up end of
the filament. Neglecting surface tension, air friction, and inertia,
the force in the filament is constant along its length, and the tensile
stress is given by:

(FE

F
vL F
= 7TR'i = Vo 7TR6

( 6-20)

Assuming that the logarithm of the filament diameter decreases


linearly with the distance from the die, the tensile strain rate at the
take up, i L> is:

( 6-21)

248

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Then the apparent extensional viscosity corresponding to 8Lis:


(6-22)
Laun and Schuch [21] compared values of this quantity with values
of the true extensional viscosity, TJE(8 L ), determined at the capillary
temperature using an extensional rheometer and found that there
was a rough correspondence.
Laun and Schuch [21] found that filament breakage always
occurred at a stress of about 0.8 MPa for all the low density
polyethylenes they studied, while the "melt strength" or force at
break varied strongly with the drawdown ratio. This implies that the
tensile stress at break is a basic physical property of the polymer
and not a rheological property. However, the drawdown required to
produce this stress is governed primarily by the extensional flow
properties.
Laun and Schuch [21] also examined the effect of die geometry
and other parameters on the reliability of melt strength measurements and made the following observations. A capillary die with a
large L / D should be used to minimize the effect of entrance flow
history and swell. A high flow rate will minimize the effects of heat
transfer on the filament temperature. A high nominal drawdown
ratio (vL/V o) and a low value of YA will minimize the effects of
swell. These guidelines are not always compatible, and some compromise is necessary in designing an experiment.
In an effort to obtain more meaningful rheological data the
filament can be extruded into a thermostatted chamber so that the
drawing is carried out isothermally. Because of the difficulty of
passing a molten filament over a wind-up drum, this element of the
apparatus is usually placed below and outside the hot chamber.
Sampers and Leblans [22] used such a technique to carry out
isothermal drawing and measured the local strain rate along the
filament. Noting that the tensile stress in the filament can often be
assumed to be constant, they compared their results with those of
an isothermal creep experiment but obtained quite different results.
They attributed this to the importance of the strain history experienced by the melt prior to leaving the die.
In an attempt to carry out a true isothermal melt strength test,
Githuku [23] used a rotary clamp as his drawdown device and

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

249

placed this inside the thermostatted oven so that the entire filament
was at a controlled, uniform temperature. While this technique
worked well for LDPE resins, other polymers often stuck to the
rotors and rendered them ineffective.
An alternative method for generating simple extension at higher
strain rates is pressure-driven converging flow. For example,
Cogswell [24, 25] has proposed a procedure for calculating an
"apparent extensional viscosity" from the entrance pressure drop
for capillary flow, and this procedure has been used in an amended
form by Shroff et al. [26]. Several assumptions are involved in this
procedure, and it is not known with certainty how the apparent
extensional viscosity that is calculated is related to well-defined
rheological properties [21]. This technique is discussed in some
detail in Section 8.5.3.
Because of the no slip condition at the wall, any pressure driven
flow in a converging channel will include a zone near the wall where
the flow is primarily a shearing deformation. In order to avoid this,
and generate a flow that is more nearly a homogeneous extensional
flow, it has been proposed that the channel walls be lubricated with
a relatively low-viscosity fluid [27-29]. However, it is very difficult to
provide a uniform layer of lubricant [29]. Furthermore, even if this
could be done, the flow would not be equivalent to simple extension, because it is not possible to reproduce simultaneously the
streamlines and the appropriate normal stress condition for simple
extension at the wall of the channel [30].
6.3.3 Experimental Observations for LOPE

An extensive series of extensional flow tests have been performed


on two low density polyethylene film resins identified as "Melt I"
[10] and "IUPAC A," the latter being one of the resins used in the
extensive IUPAC comparative study [16]. Wagner [3] has reported
that Melt I and IUPAC A are virtually identical, and we will review
results for both of these materials in this section.
Meissner [10, 11] was the first to report data for tensile start-up
flow for Melt I, and his values of the recoverable strain, 6 (6, i) at
various stages of tests carried out at several strain rates are shown
in Figure 6-12. We note that for high strain rates all the strain at
short times (small 6) is recoverable, as in the case of a crosslinked
00

250

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

3~------+-------+-------

2 1-------...,1-----

2
STRAIN,

3
E =

it

Figure 6-12. Ultimate extensional recoil (recoverable strain) following stress release during
tensile start-up flow for LDPE at several strain rates. Dashed line corresponds to complete
(rubbery) recovery. Adapted from Ref. 11. Copyright 1971 by Steinkopff Verlag. Reprinted
by permission.

rubber, while at longer times (higher 6 values) the tensile recoil


approaches a constant value that increases with the strain rate. At
very low strain rates the recoil approaches zero, suggesting a purely
viscous response.
Laun and Munstedt [9,12,31] later made extensive measurements on the similar IUPAC A resin. Their data for the tensile
stress growth function at 150a C [31] are shown in Figure 6-13. Also
shown are three times the shear stress growth function, 3Y/ +(1),
measured at a shear rate of 10- 3 s-1 (lowest curve) and the linear
viscoelastic prediction for y/;(I) based on a discrete relaxation
spectrum determined in small amplitude oscillatory shear (solid
line). The deviation of the experimental data from the linear
viscoelastic prediction at short times was attributed to the finite rise
time associated with all start-up flows. The marked extension thickening behavior of LDPE is apparent, with the deviation from linear
behavior occurring at a Hencky strain rate of about 0.5 at all strain
rates. The sharp rise in Y/;Ct, i) that occurs for the larger values of
the strain rate is in agreement with the prediction of the rubberlike

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

251

0.001
10 5

:;;;
~

''''

'"

10 4

103L-______
10

________

_______L________

_______L_____ '

10'
TIME (5)

Figure 6,13. Tensile stress growth function at various strain rates for LDPE IUPAC A at
150C. Adapted from Ref. 31. Copyright 1979 by Steinkopff Verlag. Reprinted by
permission.

liquid theory [32]. However, whereas the rubberlike liquid model


predicts that the stress increases without limit above some critical
value of the strain rate, the stress for IUPAC A (Figure 6-13)
reaches a steady state [12], and the tensile viscosity can therefore be
determined.
While the occurrence of a steady state value of 11 +(t, i) is not
apparent on the log-log plot, on a plot of log O"E versus t: there is a
zone of essentially constant stress that extends over about 2 strain
units, as shown in Figure 6-14 [12]. Using the steady state stresses,
Laun and Miinstedt [12] prepared the plot of 11E(i) for IUPAC A
at 150C that is shown in Figure 6-15. The tensile viscosity first rises
and then falls after passing through a broad maximum. At strain
rates below 10 - 3 S - \ the tensile viscosity is equal to its linear
viscoelastic value, 3110' Figure 6-16 shows the ultimate tensile recoil,
t:ooCO"E)' determined by cutting segments of the stretched samples
after the stress reaches a steady value. The ultimate recoil following
steady shear, Y(x,(O") is also shown. At low stresses, both recoils are

252

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

10 5

m
~

",'"

u)

m
w
4
a: 10
Im
w

...J

U5

I-

10 3

3
STRAIN,

4
f

= it

Figure 6-14. Tensile stress versus total strain (it) for LDPE IUPAC A at 150C with
= O.lS-I. Adapted from Ref. 12. Copyright 1978 by Steinkopff Verlag. Reprinted by
permission.

proportional to the stress, implying constant values of the ultimate


recoil functions, Roo and Sy" as predicted by the theory of linear
viscoelasticity. At higher stresses, the ultimate recoil values fall
below the linear viscoelastic predictions and appear to be approaching limiting values, although at the maximum stress achieved, using
presently available experimental techniques, a true limit has not
been observed.

Cij'

10 6

<G

~
~

""'"

10 5

VI
-

10- 3

I~~

--------31/0

10- 2
STRAIN RATE, (5-')

Figure 6-15. Tensile viscosity versus strain rate for LDPE IUPAC A at 150C. Adapted from
Ref. 12. Copyright 1978 by Steinkopff Verlag. Reprinted by permission.

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

10'

a:

~LI -

10

I-

:/

m
w
--1

III

a:

10 -,

a:

~3

~
fx

/-;7

w
>
0
U
w 10 -2

10

253

,,"

)7

10'

10'

TENSILE STRESS, (IE (Pa)


OR SHEAR STRESS, (I (Pa)

Figure 6-16. Ultimate recoil following stress release of LDPE IUPAC A at 150C as a
function of stress during steady simple extension (solid curve) and steady simple shear
(dashed curve). Adapted from Ref. 12. Copyright 1978 by Steinkopff Verlag. Reprinted by
permission.

Raible et al. [33] measured the tensile stress growth coefficient,


77;(t, i), and the recoverable strain, 8 (8), up to a Hencky strain of
00

7 and found that after broad maxima between 8 = 5 and 6, both


functions decreased at higher strains. They concluded from these
observations that there is no steady state regime for the LDPE they
studied (not IUPAC A) and that as a result, values of 77E(i) could
not be determined. At such high values of 8 it is very difficult to
maintain a uniform sample diameter, and Raible et al. used a
statistical measure of the "quality of test performance" to validate
their data. However, few other published data have included such
an evaluation of results based on many replicate experiments. More
commonly, the maximum in the curve of 77;(t, i) that occurs just as
the sample begins to neck down somewhere along its length is taken
as the steady state value.
Miinstedt and Laun [34] have carried out an extensive study of
the effects of molecular structure on the extensional flow properties
of LDPE. Two types of rheometer were used, and the tensile
viscosity from the two instruments were in agreement. It is of
interest to note that the creep test reached steady state much

254

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Table 6-1. Data for Three LOPE Samples [34]

sample
2
3

'110
X 10 4
X 10 4

3.2
3.3
3.7 X 10 4

Mw/Mn
7.0
8.2
24.2

'I1E(max)/3'110

Dg (Pa -1)

If (Pa- 1)

2.6
3.0
6.1

1.3 x 10- 4
2.0 X 10- 4
4.4 X 10- 4

2.5 X 10- 4
4.0 X 10- 4
7.5 X 10- 4

sooner than the constant strain rate (tensile start-up flow) method.
The authors could not be certain that when a decreasing stress was
observed during a tensile start-up flow, it was not due to inhomogeneities in the sample. Nevertheless, for samples with a high
degree of long chain branching, the extension was quite stable, and
there was a broad zone of constant stress, from which a value of the
tensile viscosity was calculated.
In one comparison, Miinstedt and Laun [34] looked at the effect
of molecular weight distribution on the tensile viscosity. Data for
the three LDPE resins studied are shown in Table 6-1. The three
samples all had quite similar densities, suggesting similar degrees of
long chain branching. Sample No. 3 had a distinct high molecular
weight tail, as determined by GPC measurement, and this explains
the large value of its polydispersity index.
The tensile viscosity data are shown in Figure 6-17. The presence
of the high molecular weight material has a marked effect on the
response. The authors suggest that the maximum value of 77E(6),
divided by 3770' is a useful measure of extension thickening, and we
see from both Figure 6-17 and Table 6-1 that this quantity is very
sensitive to the presence of high molecular weight material. The
steady state compliance, on the other hand, is more sensitive to a
general broadening of the MWD, as can be seen by comparing
results for samples 1 and 2. We see that the linear viscoelastic
limiting values of the steady state compliances in shear and extension do not follow the classical result given by Equation 2-44
(D2 = JP /3). This is due to the fact that the tensile stress was not
sufficiently low for true linear behavior.
In a second comparison, Miinstedt and Laun [34] looked at the
effect of molecular weight on 77E(i) for five samples thought to have
similar molecular weight distributions and degrees of branching.
Data on these resins are listed in Table 6-2. We note that the zero

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

10 6

255

p------------------------------------------,

10 3
TENSILE STRESS (Pa)

Figure 6-17. Tensile viscosity versus tensile stress at IS0C for the three LDPE materials
listed in Table 6-1. Adapted from Ref. 34. Copyright 1981 by SteinkopffVerlag. Reprinted
by permission.

shear viscosity for these highly branched materials increases much


more steeply with Mw than the 3.4 power relationship that is valid
for linear resins.
The tensile viscosities for these materials are shown in Figure
6-18, where we see that the shapes of the curves are rather similar.
The scaling with the zero shear viscosity is only approximate,
however, as is indicated by the values of 7JE(max)/37Jo shown in
Table 6-2. The authors conclude from this that the MWDs for the
five samples are not as much alike as had been thought, and in the
case of resins 7 and 8, the steady state compliance values support
this conjecture. However, GPC results did not reveal significant

Table 6-2. Data for Five LOPE Samples [34]


sample
4
5
6

7
8

Mw

805,000
687,000
467,000
327,000
245,000

7Jo (Pa s)

6.0
1.4
5.0
9.0
1.0

x
x
x
x
x

10 5
10 5
10 4
10 3
10 3

7J E(max) /37Jo

IJ(Pa- 1 )

4.0
4.5
7.0
7.6
5.3

8.0 X 10- 4
8.7 x 10- 4
5.9 x 10- 4

256

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

10 8

r------------------------------------------------,

LOPE4
3110-------------LOPE5
3110-------3110
LOPE6

LOPE?

11 0

---::::-

8
3110-----------

TENSILE STRESS,

0.

(Pa)

Figure 6-18. Tensile viscosity versus tensile stress at 150C for the five LDPE materials listed
in Table 6-2. Adapted from Ref. 34. Copyright 1981 by Steinkopff Verlag. Reprinted by
permission.

differences in MWD, and the authors concluded that the 11(i)


function is very sensitive to certain features of the molecular weight
distribution, and that differences too small to be found by chromatography can affect the tensile viscosity curves.
Finally, Miinstedt and Laun [34] studied the effect of degree of
long chain branching by comparing three resins having different
densities but similar values of 110' Data for these resins are shown
in Table 6-3. It seems clear that increased branching at constant
Mw increases the extension thickening, as indicated by the values of
11(max)/3110' while the zero shear viscosity and steady state compliance decrease.
Wagner [3] used previously published data [9,11] for 111(t, i) to
determine the tensile damping function, h(E), by use of Equation

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

257

Table 6-3. Data for Three Polyethylenes [34]

Jf

sample

Mw

(Pa s)

Mw/Mn

3110

(Pa- 1)

(g/cm 3)

CH 3 per
1000CH 2

6
9
H3

467,000
256,000
152,000

5.0 X 10 4
2.7 X 10 4
3.6 X 10 4

25
10
14

7
2.8

8.0 X 10- 4
4.2 X 10- 4
1.0 X 10- 3

0.918
0.928
0.960

30
15
0

110

lIE(max)

6-12. The results are shown in Figure 3-5. We note some depen-

dency on strain rate, suggesting that the memory function may not
be entirely separable into a time-dependent memory function and a
strain-dependent damping function, although Wagner felt the scatter in the data could be due to experimental errors. The solid curve
in Figure 3-5 is a plot of Equation 3-76 with m = 0.3 and a = e- 6
Wagner used this damping function to predict the creep and creep
recovery functions as well as the recoverable strain during start-up
of steady simple extension. The predicted behavior was in fairly
good agreement with experimental results [9, 11].
The use of Wagner's equation (3-64) to predict the recoverable
strain for simple extensional flows is not so straightforward as in the
case of simple shear flows, because the tensile recoil process involves different kinematics from the extensional flow from which it
arises [3, 35]. Laun [35] suggests the use of effective relaxation
strengths, gj(i), to calculate 8",(i). These are inferred from tensile
viscosity data by use of the following empirical equation:

( .)

8=

g,

G,
., )
( . , )2
exp(
-r8Aj + C 8Aj

(6-22)

For IUPAC A, the fitted parameters are: r = 4 and C = 0.0293.


This function has a maximum when plotted versus iAj, and this
corresponds to the maximum in the tensile viscosity curve shown in
Figure 6-15. The steady state recoverable strain is then calculated
from:
(6-23)

258

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

6.3.4 Experimental Observations for Linear Polymers

There have been reports of moderate extension thickening in linear


polymers, especially when some very high molecular weight material is present [34,36-39]. Other observations, however, indicate
that linear polymers generally exhibit strong extension thinning
[8,21,40,41]. This question is difficult to resolve for several reasons.
First, extensional flow experiments are especially difficult in the
case of polymers that are not strongly extension thickening, because
it is very important that the sample diameter be uniform throughout an experiment to avoid necking. As a result, linear materials
cannot be stretched isothermally as much as highly branched ones
[36]. This makes it particularly difficult to verify the achievement of
a steady state regime in creep and start-up experiments so that
reliable values of the tensile viscosity can be reported.
Another reason why this question of the behavior of linear
polymers is difficult to resolve is that some degree of long-chain
branching may have been present in some of the "linear" materials
that were studied. In this section we summarize some of the most
extensive reports of experimental results.
In their comparative study of 4 HDPE resins having Mw/Mn
values between 9 and 13, Miinstedt and Laun [34] found that the
two resins having the highest molecular weight (Mw "" 220,000)
exhibited a mild degree of extension thickening and a maximum in
the tensile viscosity curve at a net tensile stress between 10 3 and
10 4 Pa. However, two samples having Mw values of 98,000 and
152,000 exhibited no extension thickening and an 1](i) that was
equal to 31]0 up to (J' = 10 4 Pa, where it started to decrease
gradually. The authors concluded that the polydispersity index is
not a useful guide to the shape of the tensile viscosity curve,
because small differences in the high end of the molecular weight
distributions that have little impact on the value of Mw/Mn can
have an important effect on the extensional flow properties.
Linster and Meissner [36] measured the tensile stress growth
coefficient, 1];(t, i), of four high density polyethylenes varying in
M w ' comonomer content and molecular weight distribution. The
value of Mw varied from 150,000 to 225,000 and Mw/Mn from 7.5
to 9.4. Using the rotary clamp technique, special care was taken
with the apparatus design and experimental procedure to minimize

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

259

sample inhomogeneities and to make possible maximum elongations up to a Hencky strain of 5.6.
Extension thickening was observed in all cases, although the
effect was not as strong as in the case of LDPE. A constant stress
was achieved, but no tensile viscosity curves were shown. Short
chain branches (due to butene comonomer) did not enhance the
extension thickening, but the presence of even a small amount of
high molecular weight material had a very strong effect. Extension
thickening was more pronounced at 170C than at 150C, and this
seems to rule out the possibility that strain-induced crystallization
was causing this phenomenon.
Miinstedt [37] determined the tensile creep compliance, the tensile stress growth coefficient and the recoverable strain of four
polystyrenes having different molecular weight distributions. He
found that the behavior was dominated by the presence of even
modest amounts of high molecular weight material. The presence
of such material enhanced both the nonlinearity of the response
and the recoverable strain. In fact the polymer that had the lowest
value of M w ' but which also had a small but distinct amount of
higher molecular weight material, exhibited the most pronounced
extension thickening and the highest recoverable strain. The tensile
viscosity seemed to approach 3770 at lowe, rose to a modest
maximum and then decreased. However, it was not possible to
make a precise determination of 770. The low Mw material with a
high molecular weight fraction appeared to have the most nonNewtonian behavior, but it was also the most troublesome to test,
and reliable values of the steady state stress could not be determined.
Franck and Meissner [38] measured the tensile creep compliance
and recoverable strain for a series of polystyrenes and their blends.
They also found that even a small amount of high molecular weight
material had a strong influence on the behavior. Linster and
Meissner [39] reported that extension thickening behavior was
exhibited by the polymethyl methacrylate and polystyrene samples
they used in their tensile start-up experiments.
More recently, Laun and Schuch [21] measured the tensile viscosities of blends of two polystyrenes having quite different molecular weights. They were unable to detect any extension thickening,
even when some very high molecular weight material was present.

260

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

They concluded that significant extension thickening occurs only


when long-chain branching is present. This conclusion is supported
by earlier work [40,41].
6.4 BIAXIAL EXTENSION

The term "biaxial extension" is usually used to refer to an axisymmetric flow having kinematics described by the following velocity
field:
(6-24a)
(6-24b)
(6-24c)
where eB' the biaxial strain rate, is positive. A more precise name
for this deformation is "equibiaxial" extension. Because of the axial
symmetry, this flow can also be conveniently described using cylindrical coordinates:
(6-25a)
(6-25b)
Biaxial extension is closely related to simple extension insofar as
the velocity distribution is concerned. In fact, it is possible to define
a single extensional viscosity, 7i(i), that describes the steady state
response to simple extension for e > 0 and to biaxial extension for
e < 0 [42]. However, the two flows are kinematically quite different.
First, the relationships between IiBi) and Iz<Bi) are different, as
is shown in Figure 3-1. Secondly the types of molecular orientation
that are generated by the two flows are quite different. In simple
extension, the orientation is uniaxial and in the Xl direction,
whereas in biaxial extension it is in the plane perpendicular to the
Xl axis.
The rheological response of a material subjected to biaxial extensional flow manifests itself through the net stretching stress, aBo
(6-26)

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

261

For example, for biaxial start-up flow, with a constant strain rate
suddenly imposed at t = 0, the biaxial stress growth coefficient is:

+( t, EB
.) =_

YJB

(TB

-.-

(6-27)

EB

and the biaxial extensional viscosity is:


(6-28)
In the limit of small strain rates, linear viscoelastic behavior is
observed, and:
(6-29)
and

(6-30)
Several techniques have been used to determine the response of
polymeric liquids to biaxial extension. These include sheet inflation,
lubricated squeeze flow, diverging pressure flow and the use of
rotary clamps. Progress in this area up to 1982 has been described
by previous reviews [4-6], but there have been some interesting
developments since that time, and these are described below.
In the sheet inflation method, a circular molded sample is clamped
around its periphery, melted, and inflated by increasing the pressure in the fluid medium on one side. The deformation is not a
uniform one, with biaxial extension occurring only in the neighborhood of the center of the disk. The relationship between the
displaced volume of the inflating medium and the strain, E B , is a
complex one that cannot be predicted with any certainty. Yang and
Dealy [43] have proposed, on the basis of experiments with polyethylene, that there is a universal relationship between displacement
volume and strain near the center of the sample. This relationship
can be determined experimentally and used as the basis for a
computer algorithm to control the displacement volume as a function of time. Yang and Dealy used a liquid as the inflation medium

262

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

to avoid the complexity arising from the use of a compressible


fluid. However, if the temperature and pressure of this fluid are
monitored during an experiment, an algorithm based on a thermodynamic equation of state can be used to provide a signal proportional to displacement volume for a compressible medium. This
simplifies the design of the apparatus and the sample loading and
removal processes. Denson [44] has used such an approach in the
design of a simple, inexpensive melt tester.
In lubricated squeezing flow, a disk-shaped sample is placed
between two platens, melted, and subjected to a compressional
deformation [45). If the surfaces are not lubricated, the flow is a
complex one involving shearing and extensional modes of deformation, while if a frictionless wall contact could be provided, the
deformation would be exactly that described by Equation 6-25. Of
all the techniques that have been used to generate biaxial extension, this is the one that lends itself most readily to the generation
of step strains [46]. The problem is that as the viscosity of the
lubricating fluid is reduced, to minimize friction and thus shear
flow, the stability of the lubricant layer also decreases, leading to
dry spots and early depletion of the layer. This limits the maximum
strain that can be produced to rather small values. The situation
can be improved somewhat by the continuous pumping of lubricant
onto the disks from the center, but stability is still a problem [47].
Just as converging flow produces uniaxial extension, diverging
flow produces biaxial extension [29). As in the case of converging
flow, it is not possible to generate both the correct wall streamlines
and the correct wall stress condition using this technique [30).
However, there is a region of the flow field away from the wall in
which the flow approximates Equation 6-25 very closely, and flow
birefringence can be used to determine the stresses within this
region [48, 49].
The technique that is capable of producing the largest biaxial
strains in the closest accord with Equation 6-25 involves the use of
eight rotary clamps of the type originally developed by Meissner to
generate uniaxial extension. Meissner et al. [50,51] have demonstrated the effectiveness of this approach. This technique requires
the use of a very complex apparatus, and a high degree of skill is
required to obtain reliable results.

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

263

The use of each of the experimental methods described above


requires either elaborate equipment, exceptional skill, or both. For
this reason there has been no systematic study of the response of
molten thermoplastics to biaxial extensional flow.
6.5 PLANAR EXTENSION

Planar extension (see Figure 6-6) is a uniform deformation in which


the velocity distribution is given by:
( 6-31a)
V2

= 0

(6-31b)
(6-31c)

where i is the Hencky strain rate in the principal strain direction.


Thus there is stretching in the Xl direction, compression in the X3
direction and no displacement in the x 2 direction. This is similar to
the type of deformation that occurs in parts of the film blowing and
sheet extrusion processes.
The rate of deformation tensor has the following components:

o
o
o

( 6-32)

For a Newtonian fluid, Equation 1-49 indicates that 722 = 0 for this
flow, while Equation 2-11 shows that this is also true for linear
viscoelastic behavior. However, for a material that exhibits nonlinear viscoelasticity, this component is not zero, and for polymeric
liquids subjected to large strains it is positive. This means that to
generate a planar extensional deformation it is necessary to apply
tensile forces of different amounts in both the Xl and x 2 directions.
It also means that three rheologically meaningful normal stress

264

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

differences can be defined for this flow:


(6-33a)
(6-33b)
(6-33c)
Obviously only two of these are independent, and different sets
of two have been used in different laboratories. For example,
Meissner et al. [52] and Laun and Schuch [21] have made use of the
second and third stress differences shown above. For this choice,
two independent material functions can be defined for the transient
start-up of planar extensional flow:
(6-34)
(6-35)
The theory of linear viscoelasticity (see Equation 2-11) indicates
that for small or slow deformations:
TJp~ = 4TJ +(t)

(6-36)

2TJ+(t)

(6-37)

TJ;2

These results are only valid at very small strain rates, and for larger
values of 6, these functions will become dependent on 6 as well as
t. However, in the initial stages of the experiment linear behavior
should be observed. For this reason, it is often useful to present
results in the form of a plot showing, on a single graph, the
quantities: TJ +(t), TJp~(t, 6)/4 and TJp+z<t, 6)/2 [21,52].
The generation of planar extensional flow and the measurement
of the relevant material functions is quite difficult. The sheet
inflation technique has been used for this purpose (53,54] as well as
lubricated planar stagnation flow [55]. The most interesting results
for molten polymers have been obtained using Meissner's rotary
clamp technique [52] and a tube stretching technique developed by
Schuch [21]. The latter involves the stretching of a tube of melt in

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

265

the Xl direction while maintaining an internal pressure to prevent a


change in diameter.
Laun and Schuch [21] report that for LDPE, which shows marked
extension thickening in simple extension, the function 71p~/4 rises
sharply above the curve of 71 +(t), falling somewhat below the curve
of 71; 13 for the same value of i, which was 0.05 S-l. The second
planar extensional stress growth function does not exhibit extension
thickening, and the curve of 71;2/2 is close to that for 71 +(t).
On the basis of criteria proposed by Larson [2], planar extension
and simple shear are expected to have the same tendency to align
molecules. But based on the tendency of these flows to generate a
high degree of molecular stretching [1], steady simple shear is a
"weak" flow, while steady planar extension is a "strong" flow. As
discussed in Section 5.8, however, exponential shear is expected to
be a strong flow. These considerations suggest that there should be
a close relationship between the responses of a melt to planar
extension and exponential shear. Samurkas et al. [56] compared
exponential shear results for a LDPE with the planar extension
results of Laun and Schuch [21]. They found that the damping
functions that fitted these two flows are in fact quite different.
Thus, the role of kinematics and strain history in rheological
response appears to be more complex than these simple criteria
indicate.
6.6 OTHER EXTENSIONAL FLOWS

It is possible to define a general class of multiaxial, shear free flows

for which the rate of deformation tensor has the following components [50, 57]:

o
m

(6-38)

Meissner et al. [52] have pointed out that if i is interpreted as the


largest principal Hencky strain rate, and m is constant, then every
possible extensional flow corresponds to some value of m between
- 0.5 and + 1.0. Simple extension corresponds to m = - 0.5, m = 1

266

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

corresponds to equibiaxial extension, and m = 0 gives planar extension.


Desmarmels and Meissner [58] have used the rotary clamp technique to study the response of polyisobutylene to extensional flows
with 0 ~ m < 1. In addition [59], they have compared their results
with the predictions of several constitutive equations. They found
that while the linear theory of viscoelasticity describes the small
strain or small strain rate results satisfactorily, none of the nonlinear models they tested were able to predict the nonlinear behavior.
However, Wagner [60] found it possible to fit all their data to a
reasonable approximation by use of a generalization of Equation
3-64. He allowed the stress to depend on the Cauchy tensor as well
as the Finger tensor as follows:

T;/t)

m(t - t')h(11,!2)[(1 + f3)B ij (t,t') + f3 Cij(t,t')] dt'


00

(6-39)
The parameter f3 is equal to N 2 / Nl in viscometric flow, and
Wagner used a value of - 0.27. For the damping function, he used
the simple empirical form given by Equation 6-40, with a = 0.11:

h (I I)
l'

aJ (11 -

--r=;:====;::=;==::::::::;=-

1+

3)(12 - 3)

(6-40)

REFERENCES
1. R. I. Tanner, Engineering Rheology, Oxford University Press, Oxford, 1986.
2. R. G. Larson, Constitutive Equations for Polymer Melts and Solutions, Butterworths, Boston, 1988.
3. M. H. Wagner, 1. Non-Newt. Fl. Mech. 4:39 (1978).
4. J. M. Dealy, 1. Non-Newt. Fl. Mech. 4:9 (1978).
5. C. J. S. Petrie, Elongational Flows, Pittman, London, 1979.
6. J. M. Dealy, Rheometers for Molten Plastics, Van Nostrand Reinhold, N.Y.,
1982.
7. F. N. Cogswell, Plast. Polym. 36:109 (1968).
8. H. Miinstedt, Rheol. Acta 14:1077 (1975).
9. H. M. Laun and H. Miinstedt, Rheol. Acta 15:517 (1976).
10. J. Meissner, Rheol. Acta 8:78 (1969).

EXTENSIONAL FLOW PROPERTIES AND THEIR MEASUREMENT

11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.

267

J. Meissner, RheoZ. Acta 10:230 (1971).


H. M. Laun and H. Miinstedt, Rheol. Acta 17:415 (1978).
H. Miinstedt, J. Rheol. 23:421 (1979).
R. Muller and D. Froelich, Polymer 20:1477 (1985).
W. F. Busse, J. Polym. Sci. A-2 5:1249 (1967).
J. Meissner, Pure and Appl. Chem. 42:553 (1975).
H. H. Winter, Pure and Appl. Chem. 55:943 (1983).
J. Meissner, Trans. Soc. Rheol. 16:405 (1972).
K. F. Wissbrun, Polym. Eng. Sci. 13:342 (1973).
R. K. Bayer, Rheol. Acta 18:25 (1979).
H. M. Laun and H. Schuch, J. Rheol. 33:119 (1989).
J. Sampers and P. J. R. Leblans, J. Non-Newt. Fl. Mech. 30:325 (1988).
D. Githuku, M. Eng. Thesis, Chern. Eng., McGill Univ., Montreal, 1985.
F. N. Cogswell, Trans. Soc. Rheol. 16:383 (1972).
F. N. Cogswell, Poly. Eng. Sci. 12:64 (1972).
R. N. Shroff, L. V. Cancio and M. Shida, Trans. Soc. Rheol. 21:429 (1977).
F. N. Cogswell, J. Non-Newt. Fl. Mech. 4:23 (1978).
A. E. Everage and R. L. Ballman, Nature 273:213 (1978).
H. H. Winter, C. W. Macosko and K. E. Bennett, Rheol. Acta 18:323 (1979).
T. Hsu, P. Shirodkar, R. L. Laurence and H. H. Winter, Proc. 8th Intern.
Congr. Rheo!., 2:155 (1980).
H. Miinstedt and H. M. Laun, Rheol. Acta 18:492 (1979).
M. H. Wagner, Rheol. Acta 15:133 (1976).
T. Raible, A. Desmarmels and J. Meissner, Polym. Bull. 1:397 (1979).
H. Miinstedt and H. M. Laun, Rheol. Acta 20:211 (1981).
H. M. Laun, J. Rheol. 30:459 (1986).
J. J. Linster and J. Meissner, Polymer Bull. 16:187, (1986).
H. Miinstedt, J. Rheol. 24:847 (1980).
A. Franck and J. Meissner, Rheol. Acta 23:117 (1984).
J. J. Linster and J. Meissner, Makromol Chem. 190:599 (1989).
Y. Ide and J. L. White, J. Appl. Polym. Sci. 22:1061 (1978).
F. N. Cogswell, Trans. Soc. Rheol. 16:383 (1972).
J. M. Dealy, J. Rheol. 28:181 (1984).
M.-C. Yang and J. M. Dealy, J. Rheol. 31:113 (1987).
C. D. Denson, "Costech 2000 Process Simulator", Costech Associates Inc.,
1184 Corner Ketch Rd., Newark, DE 19711.
Sh. Chatraei, C. W. Macosko and H. H. Winter, J. Rheol. 25:433 (1981).
P. R. Soskey and H. H. Winter, J. Rheol. 29:493 (1985).
A. C. Papanastasiou, C. W. Macosko and L. E. Scriven, in Interrelations
between Processing Structure and Properties, edited by J. C. Seferis and P. S.
Theocaris, Elsevier Scientific Publishers, Amsterdam, 1984.
J. A. Van Aken and H. Janeschitz-Kriegl, RheoZ. Acta 19:744 (1980).
J. A. Van Aken and H. Janeschitz-Kriegl, Rheol. Acta 19:744 (1980); 20:419
(1981).
J. Meissner, T. Raible and S. E. Stephenson, J. Rheol. 25:1 (1981).
J. Meissner, Polym. Eng. Sci. 27:537 (1987).

268

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

52. J. Meissner, S. E. Stephenson, A. Desmarmels and P. Portman, 1. Non-Newt.


Fl. Mech. 11:221 (1982).
53. C. D. Denson and D. L. Crady, 1. Appl. Polym. Sci. 18:1611 (1974).
54. C. D. Denson and D. C. Hilton, Polym. Eng. Sci. 20:535 (1980).
55. R. B. Secor, C. W. Macosko and L. E. Scriven, 1. Non-Newt. Fl. Mech. 23:355
(1987).
56. T. Samurkas, R. G. Larson and J. M. Dealy, 1. Rheol. 33:559 (1989).
57. J. F. Stevenson, S. c.-K. Chung and J. T. Jenkins, Trans. Soc. Rheol. 19:397
(1975).
58. A. Desmarmels and J. Meissner, Rheol. Acta 24:253 (1985).
59. A. Desmarmels and J. Meissner, Coli. Polym. Sci. 264:829 (1986).
60. M. H. Wagner, 61st Annual Mtg., Soc. of Rheology, Montreal, 1989.

Chapter 7
Rotational and Sliding
Surface Rheometers

7.1 INTRODUCTION

There are two basic types of instrument for measuring shear properties: capillary and slit rheometers in which the flow is generated
by a pressure drop; and drag flow rheometers in which one bounding wall moves relative to a second, stationary wall. Pressure-driven
rheometers are described in Chapter 8, and the present chapter
deals with drag flow instruments.
In order to generate results that are subject to interpretation in
terms of well-defined material functions, it is desirable that the flow
generated by the rheometer be as close as possible to simple shear
flow. This can be accomplished with a minimum of mechanical
complexity by the use of either rotational or rectilinear motion of a
solid surface. Drag flow rheometers based on rectilinear motion are
referred to here as "sliding surface rheometers." Both sliding plate
and sliding cylinder rheometers are described in this chapter.
While pressure flow rheometers are popular because of their
simplicity and ease of use, they provide information only about the
shear stress associated with steady shear. Thus, they are used
primarily for the measurement of viscosity at high shear rates.
Furthermore, since the shear rate is not uniform, it is not possible
to obtain the true value of the viscosity at one shear rate from a
single experiment. As is explained in Chapter 8 a somewhat elaborate treatment of data obtained at several flow rates is necessary to
obtain the viscosity curve. With the possible exception of the first
normal stress difference, reliable information about the viscoelastic
269

270

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

behavior of a material cannot be obtained by use of a pressure-flow


rheometer, because uniform transient flows cannot be generated.
For these reasons, studies of melt viscoelasticity require the use
of rheometers capable of generating a uniform deformation that
can be precisely controlled as a function of time. Extensional flow
rheometers capable of accomplishing this objective are described in
Chapter 6. While for studies of nonlinear viscoelasticity, extensional
and shear rheometers each reveal different aspects of the rheological nature of a material, either could, in principle, be used to
determine the linear properties. However, because they are much
easier to build and use, drag flow shear rheometers are always used
to determine the linear viscoelastic properties of molten polymers.
Of course, drag flow shear rheometers have their own limitations
and sources of error, and these are summarized in the present
chapter. More complete treatments of rotational and sliding surface
rheometers are given by Dealy [1] and in the book edited by Collyer
and Clegg [2, Chapters 9, 12, 13].
7.2 SOURCES OF ERROR FOR DRAG FLOW RHEOMETERS

While each type of drag flow rheometer has its specific limitations,
certain problems are common to all such instruments, and these are
described in the present section.
7.2.1 Instrument Compliance

If the only relative motion of the confining walls of a drag flow

rheometer is the displacement imposed to generate the desired


shear deformation, then the shear strain, y, in the sample is directly
related to this imposed motion. However, if the instrument is
compliant, i.e., if it deforms in response to the forces exerted on the
test fixtures by the sample, then the relationship between the shear
strain and the imposed displacement is more complex. Moreover,
this relationship depends on the rheological properties of the melt,
and it therefore varies with time during a transient shear experiment.
This does not pose a problem for a steady shear experiment such
as the measurement of viscosity. However, for the measurement of
transient properties such as the relaxation modulus or the storage

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

271

and loss moduli it can lead to significant errors. In addition, it can


cause difficulty in the measurement of the first normal stress
difference in a rotational rheometer, even in the case of steady
shear. This latter effect will be examined in more detail in the
section on cone-plate rheometers, and we consider here only the
effect of instrument compliance on the measurement of shear stress
in a transient flow.
Instrument compliance results from the deformation of the
rheometer frame and of the transducer used to sense the shear
stress. In the case of a rotational rheometer, the latter will be a
torque transducer, while in the case of a sliding surface rheometer
it will be a force transducer, often called a load cell. As a result of
this compliance, the actual deformation in the sample will be less
than the apparent value calculated from the imposed displacement.
In addition, due to the viscoelastic nature of the sample, there will
be a time lag between the imposition of the overall displacement
and the actual deformation of the sample. Lockyer and Walters [3]
have shown how the use of a compliant torque transducer in a
rotational rheometer can produce an apparent shear stress overshoot at the start-up of steady shear, even in a Newtonian fluid.
To understand this phenomenon more quantitatively, it is useful
to refer to the simplified model shown in Figure 7-1. The instrument frame and transducer are modelled as linear (Hookean)
springs, while the fluid is modelled by one Maxwell element. For a
sliding surface rheometer, in which a linear actuator generates the
deformation, and the shear stress is inferred from the output of a
force transducer, this force corresponds to the force "F" in the
figure. For a rotational rheometer in which a torque transducer is
used, this force is analogous to the torque.
In Figure 7-1, the displacement of the actuator is dX, while the
instrument is stretched or compressed by an amount dY due to the
compliance of the frame and the transducer. Thus the deformation
of the sample is not dX, but dX - dY. Furthermore, the ratio
(d X - d Y) / d X depends on the properties of the sample and is
not constant during a transient experiment. Thus, instrument
compliance cannot be compensated for by a simple calibration
procedure.
.
For the model described above, if inertia is neglected the force is
equal in all three elements. In this case, the apparent values of the

272

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

INSTRUMENT FRAME

TRANSDUCER

FLUID

Figure 7-1. Simple model for the effect of instrument compliance. The instrument frame and
transducer are modelled as linear springs, while the fluid is modelled as a single Maxwell
element.

storage and loss moduli that would be inferred from the total
displacement in a small amplitude oscillatory shear experiment
would be as follows:

G'=
a

(7-1)

(7-2)

where:

= viscosity
A = relaxation time of fluid
c = instrument compliance

TJ

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

273

The constant K depends on the geometry of the shear fixtures. For


example, for cone-plate geometry, K = 30/27rR 3 , where 0 is the
cone angle.
The time constant for the instrument compliance effect is the
quantity (CT/ / K), and to minimize the resulting error, this quantity
must be much smaller than A. In a step strain experiment, this same
time constant appears in an exponential term that causes the
deformation of the sample to lag behind the ideal step function.
Clearly, the high viscosity of molten polymers will enhance this
effect.
An alternative to the redesign of the instrument to reduce the
time constant is to correct experimental data to account for compliance. For the case of small amplitude oscillatory shear this is
straightforward, and Laun [4] gives the following formula relating
the true moduli to their apparent values:
(7-3)

where k is the modulus of the force transducer.


Obviously it is advantageous to make the instrument compliance
as small as possible to eliminate the problem entirely, and the
stiffening of the rheometer frame will improve the situation [5, 6].
However, there is a limit to the extent that the transducer compliance can be reduced. This is because force and torque transducers
are all based on the sensing of the deformation of the load
transmitting element of the transducer in response to the force or
torque. Thus, as the transducer compliance is reduced, the magnitude of its output signal is reduced as well as its signal to noise
ratio.
One way around this problem is to use a servo positioning loop to
apply a compensating displacement of a second compliant member
such that the displacement of the rheometer fixtures is maintained
at zero. The deflection of the second compliant member is now
used to track the torque or force. However, in this case a new
instrument time constant is introduced, the one associated with the
control loop. Another approach to the problem is to use a piezo-

274

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

electric transducer, which has a very low compliance [6,7]. However, this type of sensor operates on the basis of the buildup of
electric charge due to the deformation of the sensor, and this
charge will inevitably leak away with time, decaying more rapidly
the higher the temperature. Therefore, this approach is useful only
to track the force during short transient experiments or for relatively high frequency oscillatory shear. The rate of charge decay can
be minimized by the use of an electrometer to monitor the charge
directly [6].
7.2.2 Viscous Heating

In presenting the results of rheological measurements, it is almost


always assumed that the sample is at a uniform, known temperature. However, this can never be precisely the case because of
viscous dissipation. Whenever a viscous material is deformed, some
of the work of deformation is converted to internal, thermal energy
by means of viscous dissipation. This increase in thermal energy will
result in a temperature increase, unless it is removed by the flow of
heat. However, heat only flows when there is a temperature gradient. Thus, a temperature gradient in the sample is inevitable. To
illustrate this point, consider the steady-state sliding plate shear of
a Newtonian fluid with a constant viscosity. If both plates are at a
fixed temperature, To, the steady state temperature distribution is
as follows:

(7-4)
The maximum temperature rise occurs at the midplane and is:

(7-5)
Clearly, viscous heating will be a problem when T7, y, and hare
large and k is small. However, the use of the steady state analysis
provides a worst case estimate of the maximum temperature rise,
and in practice we can often make our rheological measurement

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

275

before sufficient viscous heat has been generated for the temperature to approach its steady state distribution.
A detailed analysis of the viscous heating problem is given by
Warren [2, Chap. 5].
7.2.3 End and Edge Effects

Because the sample is of finite size, in drag flow rheometers there is


usually an interface between the edge of the sample and the
surrounding medium, which is normally air or nitrogen. This interface is not taken into account in the basic equations for calculating
stress and strain, and the stress and deformation fields in the
neighborhood of such a free surface deviate from those assumed in
deriving these equations. In some cases edge or end effects can be
so severe that they make it impossible to carry out tests under
certain conditions. In cone-plate and parallel disk rheometers, one
observes severe distortion of the free surface, while in concentric
cylinder instruments, rod climbing (the Weissenberg effect) interferes with the measurement. Edge and end effects can also cause
errors in results from sliding surface rheometers, unless a shear
stress transducer is used. These problems will be discussed in
further detail in the following sections on each type of rheometer.
7.2.4 Shear Wave Propagation

Another source of inhomogeneous deformation is shear wave propagation. This is only a factor in nonsteady test modes, but it is
precisely this type of test that is necessary to study viscoelasticity.
Shear wave propagation is a potential problem in any rheometer
when transient tests are to be performed, but it is most prominent
in experiments involving large strains and strain rates.
Because of inertia in the sample material, a change in the
velocity of the moving surface cannot instantaneously cause the
acceleration of every material element that would be necessary to
produce uniform strain in the sample. In a Newtonian fluid subjected to the sudden start-up of motion by a plate moving at
constant velocity, the velocity distribution is given by a Fourier
series [8]. At longer times, the series is dominated by the first term,
and the asymptotic approach to steady state becomes exponential,

276

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

with a "half life" of h 2p/(7r 2 TJ). We note that the approach to


steady state is most rapid when the gap is small and the viscosity is
large. This phenomenon is thus only a serious problem for liquids
of rather low viscosity, in which case it is necessary to use the
smallest possible gap.
The behavior of elastic liquids at the inception of steady shear
flow has been analyzed theoretically [9,10] and observed experimentally [11]. One possible effect of sample inertia is an overshoot
in the measured stress. Since stress overshoot is a common feature
of the stress growth function in viscoelastic fluids, care must be
taken to ensure that an observed overshoot is a rheological effect
rather than an inertial effect.
It is of interest to consider in some detail the case of oscillatory
shear, which is widely used to study both linear and nonlinear
viscoelasticity. The strain is given by Equation 2-45:

yosin(wt)

(2-45)

For a Newtonian fluid, the criterion for neglecting the effect of


inertia on the strain field is that:

(7-6)
MacDonald et al. [12] considered the case of small deviations
from a linear velocity profile in deriving a criterion for neglect of
inertial effects. In order for the storage modulus to be reasonably
free of errors due to acceleration, their criterion is:

(7-7)
while for negligible error in the loss modulus, the criterion is:
(7-8)

Schrag [13] has carried out a more thorough analysis, and his

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

277

general criterion for neglecting the nonlinearity of the displacement


profile is:

h/As 1

(7-9)

where As is the shear wavelength, which is related to the complex


modulus and the loss angle as follows:
As

27T

= --====-----

w"; p/Gdcos(8/2)

(7-10)

For the case of a fluid having a density of 1 gm/cc and a complex


viscosity of 1 Pa' s, Schrag calculates that the use of frequencies up
to several Hertz requires that h/As be less than 1/30.
7.3 CONE-PLATE RHEOMETERS

The most popular rheometer for the study of the viscoelastic


properties of molten plastics is based on the use of a circular disk
and a small-angle cone. These two elements are mounted on a
common axis of symmetry as shown in Figure 7-2, the sample is
inserted between them, and one element is rotated while the other
is held stationary. Usually the motion is programmed, and the

Figure 7-2. Cone-plate rheometer fixtures.

278

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

resulting torque, M, is measured. The reverse technique, fixing the


torque and measuring the resulting rotational displacement, is also
possible. This latter technique is useful for the measurement of the
creep compliance, the recoverable shear, and the yield stress, if
there is one. The normal (axial) force, F, exerted by the melt on the
fixtures is related to the first normal stress difference, N l
The reasons for the popularity of the cone-plate flow geometry
for use with melts are as follows:
1. Only a small sample of material is needed.
2. The shear rate is approximately uniform throughout the
sample.
3. There is ready access to all surfaces to facilitate sample
loading and cleaning.
The use of this geometry for highly elastic melts, however, is
limited because of serious departures from the assumed simple
shear flow that occur above a certain level of shear stress. This
limits its use to rather low shear rates. While cone-plate geometry is
suitable, in principle, for the measurement of the first normal stress
difference, in practice there are many pitfalls associated with the
difficulty of setting and maintaining a precise spacing between the
cone and the plate. For these reasons, cone-plate rheometers have
proven useful mainly for the measurement of the low-shear rate
viscosity and the linear viscoelastic properties of melts.
7.3.1 Basic Equations for Cone-Plate Rheometers

When certain simplifying assumptions are valid, the equations of


motion can be used to derive quite simple equations for calculating
quantities of rheological importance from cone-plate data. These
assumptions are as follows:
1. The inertia (acceleration) terms in the equations of motion
can be neglected.
2. The cone angle is sufficiently small that certain trigonometric
identities can be used to simplify the equations.
3. The surface tension acting at the exposed edge surface has no
effect on the torque or normal force.

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

279

4. The free surface at the edge of the gap between the cone and
the plate is spherical in shape with a radius of curvature equal
to the cone radius, and the flow is uniform out to this surface.
Assumptions 1 and 3 are generally valid for molten, high polymers, and assumption 2 does not lead to significant errors for the
range of cone angles normally used in practice. Assumption 4,
however, is valid only under certain circumstances. This limitation
and other sources of error for cone-plate rheometers are described
in the next section.
The measurable quantities and fixed parameters for cone-plate
flow are the fixture radius, R, the cone angle, 9 0 , the rotational
speed, n (rad/sec), the torque, M, and the normal force, F. When
the above assumptions are valid, the following equations can be
used to calculate quantities of rheological importance.
(7-11)

(7-12)
(7-13)

Equation 7-13 neglects the contribution to the normal force due


to centripetal acceleration, but for melts within the range of
usable shear rates in a cone-plate rheometer, this contribution is
negligible.
7.3.2 Sources of Error for Cone-Plate Rheometers

Secondary flow results from the radial pressure gradient associated


with centripetal acceleration and is present at all speeds [14].
This motion represents a deviation from the simple shear flow
assumed in deriving Equation 7-11, and is thus a source of error. In
Newtonian fluids, its severity is governed by a Reynolds number
[15]. For non-Newtonian, inelastic fluids, the strength of the secondary flow is enhanced by shear thinning, but for molten polymers,
because of the very high viscosity, the resulting error is negligible.
Another cause of shear rate variation in the sample is the use of a
nonzero cone angle, as the derivation of Equation 7-11 assumes this

280

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 7-3. Photo of the free surface at the edge of a cone-plate rheometer showing irregular
flow. Reproduced from Ref. 18. Copyright 1986 by The Society of Rheology. Reprinted by
permission of John Wiley & Sons, Inc.

angle to be very small. For cone angles less than 8, however, this
variation is probably less than 5%.
Viscous heating inevitably produces a temperature gradient in
the sample. However, for other reasons the shear rate is limited to
low values, and a small cone angle is used. As a result viscous
heating does not usually result in a serious temperature inhomogeneity. However, temperature gradients and variations with time
can be caused by an inadequate temperature control system. It is
thus essential to take great care in the design and operation of
the temperature control system, especially when one wants to
measure N 1
The primary source of error in cone-plate rheometers in the case
of elastic melts is the serious flow irregularity sometimes called
edge fracture [16-18]. It is usually associated with an obvious
distortion of the free surface, as shown in Figure 7-3. Gleissle
[19,20] has observed the development of this distortion during the
start up of steady shear by using a transparent plate and a silicone
polymer. He observed that the disturbance grows rapidly inward
and results in a significant decrease in the effective radius of the

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

281

sample. Nazem and Hansen [21] found evidence of a pronounced


nonuniformity in the shear strain across the gap near the outer
edge at shear rates well below those at which edge irregularities
become readily visible.
It has been observed that this problem often occurs at shear
stresses around 5 X 10 4 Pa. However, it is thought to be a manifestation of the elasticity of the melt, and Hutton [22] has suggested
that a more general measure of its severity is the quantity N 1R0 o.
The problem has been treated theoretically by Tanner and Keentok
[23] using linear elastic fracture mechanics, and the resulting criterion for fracture is:

2f

IN21> 3;
where:

H
K1

=
=

4f
=

3HK1

(7-14)

surface tension of the polymer


crack size
rheometer gap at edge of fixtures (R0)
empirical constant of order one

Using data of Macosko and Morse [24], Tanner and Keentok [23]
concluded that for LDPE IUPAC A at 150C the critical value of
the second normal stress difference is 0.5 kPa, leading to shear
stress errors at a shear rate of 0.07 s -1 and visible edge fracture
at 1 S-l.
7.3.3 Measurement of the First Normal Stress Difference

For several decades there has been a keen interest in the measurement of the first normal stress difference of elastic liquids, and
there are several reasons for this. In Chapter 4 it is shown that the
first normal stress difference is a very strong function of molecular
weight, especially at low shear rates. Furthermore, transient normal
stress results have been found useful in the evaluation of constitutive equations [25,26]. Finally, for many years it was the only
nonlinear shear property that could in principle be measured using

282

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

commercially available instruments. For these reasons, there has


been considerable interest in the use of cone-plate rheometers to
measure N J in melts.
However, many years of experience have shown that such a
measurement is subject to sources of error not encountered in the
measurement of the shear stress [25]. In particular, the normal
force, F, from which N J is calculated, is much more sensitive to
variations in gap spacing and sample temperature (See Equation
4-35) than the torque. Thus, instrument compliance and temperature fluctuations must be reduced to very low values in order to
obtain reliable values of N J
Variations in the gap spacing obviously will cause an error simply
because of the deviation of the test geometry from that assumed in
deriving Equation 7-13. However, if the gap varies during the
experiment, there will be a contribution to the normal force due to
the squeeze flow of the sample, e.g., if the gap increases, the force
will be reduced. Variations in the gap can result from bearing
run-out in the drive motor or from instrument compliance. Bearing
run-out will produce a periodic variation of the plate spacing [27],
while instrument compliance causes a variation in the gap in
response to the normal force generated by the fluid. Gleissle
[19,20,28] has measured the effect of compliance-generated squeeze
flow on the pressure distribution at the cone surface. This pressure
is related to the value of N J that would be inferred from the normal
force, F, and it was found that during start-up flow the error due to
squeeze flow is quite large at short times.
Unlike other sources of error for cone plate rheometers, this
effect increases as the cone angle is reduced, and it has been found
necessary to use cone angles greater than 4 to obtain consistent
results for transient normal force results [5, 29]. While instrument
compliance can be reduced by stiffening the rheometer frame [5],
the force transducer must have some compliance in order to function. For steady state measurements, a servo-positioner can be used
to maintain the gap spacing at its nominal value. However, for
transient measurements the servo loop will introduce its own signal
delay [5].
Zapas et al. [30] have reported that the error in transient normal
force results due to instrument compliance is greatly amplified by
the "constrained cylinder" effect. The sample adheres to the cone

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

283

and the plate over a circular zone having a certain radius, and this
radius does not change during an experiment. Thus in compression
the sample will bulge slightly, while in tension it will neck slightly.
Finally, if there is any change of the sample volume during an
experiment, it will result in some combination of squeeze flow and
increased normal force. A change in sample volume will result from
any temperature variation, which can be due to the imperfect
operation of the temperature control system or to viscous heating
[30]. In addition, a change in sample volume could be caused by a
coupling between the bulk and shear modes, with a change in
volume caused by the shear stress, although such an effect has not
been observed to date [31].
The Rheometrics RMS800 rheometer is an example of a commercial instrument designed expressly for normal force measurement. Instrument frame compliance is very low, and a "Force
Rebalance Transducer" is available that incorporates a high speed
servo positioner to prevent the axial motion of the plate due to
transducer compliance. Meissner et al. [32] have modified this
instrument by adding an elaborate temperature control system to
minimize temperature variations. They also used a specially designed plate that was comprised of two sections, with the torque
and normal force transducer coupled only to the inner section. By
varying the radius of this section it was possible to determine the
ratio N 2/N1 For an LDPE at 150C they found that N2 = -0.34N1
7.4 PARALLEL DISK RHEOMETERS

In terms of flow regularity [33] and ease of sample preparation, the


parallel disk (or "parallel plate") geometry has advantages over the
cone-plate geometry. On the other hand, the shear strain is not
uniform in the sample, and this complicates the study of materials
exhibiting nonlinear viscoelasticity, but special techniques have been
developed to estimate nonlinear properties from parallel disk data.
These include methods for determining the viscosity [34, 35], the
stress growth and stress relaxation functions associated with steady
shear [36], the nonlinear relaxation modulus [37] and other properties [38-41]. However, at the present time, the parallel disk geometry is used primarily for the measurement of linear viscoelastic
properties. The equations for calculating the storage and loss

284

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

moduli are as follows:

G' =

2Moh
4

7TR cPo

(7-15)

cos ()

(7-16)

7.5 ECCENTRIC ROTATING DISKS

The eccentric rotating disk geometry (Maxwell orthogonal rheometer) has been used to determine linear viscoelastic properties
[42-45]. The arrangement of the fixtures is shown in Figure 7-4.
Two parallel disks are arranged with an offset, d, between their
parallel axes. One is rotated at an angular speed, fl, while the
other, with its shaft supported by an air bearing, is assumed to
rotate at the same speed. The forces in the x and y directions are
measured, and the equations for calculating the 7]' and G' are as

Figure 7-4. Arrangements of fixtures in eccentric rotating disk rheometer. Ideally, both disks
rotate at the same speed.

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

285

follows:

(7-17)
(7-18)
The frequency, lV, is equal to the angular speed, .0., and the strain
amplitude, 'Yo, is equal to d/h, where h is the gap between the
disks. As in all experimental methods there are several sources of
error [1]; those resulting from instrument compliance have been
discussed by Gottlieb and Macosko [46].
It has been proposed that Fz is simply related to the normal
stress differences, but Goldstein and Schowalter [47] have shown
that this is not the case.
Before the advent of the use of microcomputers for rheometer
control and data acquisition, this type of rheometer provided the
most convenient way to measure these properties, and it is especially advantageous for studies at very low frequencies. However, at
the present time, oscillatory shear using cone-plate fixtures is the
favored technique.
Eccentric rotating disks have not been found to be generally
useful for the study of nonlinear effects, as the results cannot be
related quantitatively to nonlinear material functions [48].
7.6 CONCENTRIC CYLINDER RHEOMETERS

Concentric cylinders are, theoretically, the best geometry for a


rotational rheometer, as the ideal flow is an exact solution of the
entire equations of motion [49]. Furthermore, when the gap is
small, the shear rate becomes practically uniform [50]. In fact, such
rheometers have been used for the study of nonlinear viscoelasticity
in polymeric liquids [51-54]. However, the Weissenberg (rod climbing) effect tends to draw fluid out of the gap [55]. In addition,
loading and cleaning is inconvenient, and maintaining concentricity
is difficult when small gaps are used, especially if the rheometer is

286

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

operated at elevated temperature. For these reasons, concentric


cylinder rheometers are not often used for molten polymers.
An exception, however, is the melt elasticity indexer developed
by Maxwell [56]. This instrument was designed to be used for the
routine evaluation of resins on the basis of elastic recoil following
steady shear. Because it was designed to be simple and inexpensive,
the temperature control system, sample geometry and rotor support
system are not appropriate for scientific measurements. A commercial version of the melt elasticity indexer is sold by Custom
Scientific Instruments, Inc.
Further information on the concentric cylinder test geometry can
be found elsewhere [1].
7.7 CONTROLLED STRESS ROTATIONAL RHEOMETERS

Rheometers in which the controlled variable is the strain or strain


rate are not suitable for the measurement of creep, recovery and
yield stress. One approach to this problem is to use a servo loop
with a controlled strain rheometer to maintain the stress at a fixed
value, but for tracking long term deformation at low stress levels, it
is highly advantageous to control the stress directly.
This can be accomplished by suspending the rotating member by
means of a low friction device and applying the desired torque by
means of a drag cup motor. This is the operating principle of the
torsional creep and recovery apparatuses designed by Plazek [57]
and by Link and Schwarzl [58]. In these rheometers, the rotor is
suspended by means of a magnetic field, and very precise measurements can be made over an exceptional range of torques.
Even if friction in the rotor support system is eliminated entirely,
there will still be some inertial resistance to the acceleration or
deceleration of the rotor. Jones et al. [59] have studied the inertial
resistance of the fluid and found that while it can be significant in
the concentric cylinder geometry, it is minimal for the cone-plate or
parallel disk geometries. All of the controlled stress rheometers
mentioned in this section make use of parallel disk geometry.
A commercial version of the instrument developed by Plazek, the
MECA Creep Rheometer is available from Time-Temperature
Instruments, Inc. Another commercial controlled stress melt rheometer is the Rheometrics Stress Rheometer [60,61]. This instru-

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

287

ment makes use of an air bearing rather than a magnetic suspension system and therefore has a relatively restricted torque range.
7.8 TORQUE RHEOMETERS

While they are not truly rheometers, "torque rheometers" are


widely used for routine evaluation and quality control. Dealy [1] has
described these instruments in some detail. The basic components
are a programmable motor drive system, a torque sensor and a
mixing head. The mixing head contains two interconnecting cylindrical chambers in which the melt is agitated by a set of matched
mixing blades. The torque required to rotate the blades provides a
measure of the consistency of the melt. However, neither the
temperature nor the strain rate are uniform within the sample, and
for this reason torque rheometers are not suitable for the determination of well-defined rheological properties.
Torque rheometers suitable for the evaluation of molten plastics
are offered by Haake Buchler Instruments and by C. W. Brabender.
7.9 SLIDING PLATE RHEOMETERS

The generation of shear deformations by the linear motion of one


flat plate relative to another has certain advantages over the use of
rotational flows. Edge failure is a much less severe problem so that
higher shear rates can be reached for elastic liquids. Anisotropic
materials can be studied by variation of the angle between the
sample orientation and the direction of motion, and versatile tensile
test frames and actuators can be used to support and drive the
rheometer for high-force, large-displacement applications.
However, in contrast to the situation with rotational rheometers,
the total strain is limited by the rheometer length, and to make
possible sufficient total strain for studies of large nonlinear effects,
the test fixtures must be larger than those used in rotational
rheometers. In order to provide the maximum shear strain for a
given plate displacement, and to minimize errors due to shear wave
propagation and viscous heating, it is desirable to make the gap as
small as possible. This means that the flatness and parallelism of
the plates must be maintained to very close tolerances. Also, any
fluctuation in the gap spacing during operation must be kept very

288

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

small. Thus, the machining and alignment of the plates must be


carried out with great care. Furthermore, the mechanism used to
support and guide the moving plate must have minimal play, i.e., it
is necessary to have a very tight fit of the sliding or rolling components of the support mechanism for the moving plate. However, any
friction arising from the motion of these components will contribute
to the measured force, F, and will result in an error in the
calculation of the shear stress.
7.9.1 Basic Equations for Sliding Plate Rheometers

The basic features of the sliding plate geometry are shown in Figure
7-5. Sliding plate rheometers are either strain-controlled or stresscontrolled. Most generate a controlled rate of deformation and use
a load cell to measure the total shearing force. The deformation
can be generated by the crosshead of a tensile testing machine, or
by some other electromechanical or servohydraulic linear actuator.
Conversely, the force can be imposed by a suspended weight or by a
feedback loop containing a servohydraulic actuator, and the resulting strain can then be measured.
The shear stress, CT, is determined by measuring the force, F,
required to drive the motion of the moving plate (or the force
required to hold the stationary plate in place) and dividing it by the
wetted area of the plates, A.
CT =

The shear strain, -y,

IS

(7-19)

F jA

the displacement of the moving plate, X,

SLIDING PLATE

FLUID

I--~F

Figure 7-5. Basic features of sliding plate geometry.

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

289

divided by the distance between the plates, h:


'Y

The shear rate,


the gap:

X/h

y, is the velocity, V,

(7-20)

of the moving plate, divided by

y = V/h

(7-21)

7.9.2 End and Edge Effects in Sliding Plate Rheometers

When elastic materials are to be studied at large shear strains,


errors due to end/edge effects arise from two causes; one of these
is surface tension, which contributes to the measured force, and
the second is nonhomogeneous flow near the free surfaces of the
sample.
Laun and Meissner [62] have estimated the error due to surface
tension in the shear stress calculated using Equation 7-19. They
assumed that the sample deforms in such a way that the free
surfaces are always flat, i.e., that the sample is always a parallelepiped. For their low shear rate studies of molten plastics, they
found that this contribution was negligible. However, surface tension did pose a problem in creep recovery (recoil) experiments
where the intention is to reduce the actual shear stress to zero. As
is explained below, the free surfaces cannot behave in the simple
way assumed by Laun and Meissner, because this would require an
imbalance of the forces acting at the free surfaces. However, it is
likely that these effects themselves contribute larger errors to the
calculated shear stress than the surface tension.
We turn now to the question of nonhomogeneous flow near the
free surfaces of the sample. This results from the fact that the stress
field in a material whose ends behave in the ideal way shown in
Figure 7-4 cannot possibly balance that which exists in the surrounding medium. This medium is usually air, but the mismatch
problem persists as long as it has rheological properties different from those of the sample, even if the only difference is that
the sample and the surrounding medium are Newtonian fluids
with different viscosities. This problem was first pointed out by
Philippoff [63].

290

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

In order to maintain a balance between the stresses in the sample


and in the surrounding medium, the free surfaces will be distorted.
Another possible result is a tearing of the sample (cohesive failure)
or a pulling away from the wall (adhesive failure).
The problems resulting from end effects in sliding plate flow are
most severe when elastic materials are subjected to large strains.
Theoretical treatments of end effects in elastic materials [64,65]
predict that the apparent shear stress calculated on the basis of
Equation 7-19, with the sample assumed to be a parallelepiped, is
below the true shear stress and decreases as the length to thickness
ratio, Ljh, increases.
Edge irregularities have been observed in viscous gels [66], in
rubber sheets [67], elastomers [68], and molten plastics [69,70]. It is
clear that the importance of end effects is strongly dependent on
the nature of the material being studied and the maximum strain
and strain rate involved.
7.9.3 Sliding Plate Melt Rheometers

Sliding plate rheometers have been used for studying nonlinear


viscoelasticity of concentrated polymer solutions [71-74], bulk
liquid polymers [75-80], molten plastics [81-83], and filled polymers
[84]. Commercial instruments designed to measure the linear viscoelastic properties of solid materials, such as the Rheovibron, the
Metravib Viscoelasticimeter and the Dynastat, can be modified for
use with viscoelastic liquids by use of sandwich or sliding cylinder
fixtures [85,86].
In order to avoid the problems associated with mechanisms
designed to maintain the spacing of the plates, many of the sliding
plate rheometers that have been used have been of the "sandwich"
type illustrated in Figure 7-6. There are two sample layers and two
outer plates, with a central plate. Here there are no net lateral
forces on the central plate, and gap uniformity is maintained by the
samples themselves. This requires, however, that the moving and
sliding plates be precisely aligned and that the distance between the
outer plates be uniform and unaffected by forces arising from the
test. Furthermore, both samples must have the same size and shape
and must be positioned precisely opposite each other on either side
of the central plate. In order to avoid asymmetrical gravitational

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

291

Figure 7-6. Basic elements of a "sandwich" rheometer. The use of two sample layers
balances the lateral forces on the upper and lower fixtures due to normal stress differences.

forces on the central plate, the rheometer is usually mounted


vertically.
Weight-driven sandwich rheometers have been used to determine
the low shear rate viscosity and steady state compliance of thermoplastics [62]. Laun [81] has developed another sandwich rheometer
in which the middle plate is fixed and the outer plates are rigidly
mounted on a moving frame driven by a pneumatic cylinder. This
instrument is thus basically a creepmeter; it was specifically designed to study molten plastics at shear stresses approaching those
occurring in melt processing operations.
Soong and his coworkers [72, 73] used a sandwich type sliding
plate rheometer for the study of nonlinear viscoelasticity in concentrated polymer solutions. In order to maintain, alignment, it was
found necessary to use guide rods for the outer plates. Friction in
the bearings resulted in a noisy stress signal from the load cell, and
the studies were limited to shear rates less than 5 s - 1. A wide
variety of strain histories were generated by the use of programmable servohydraulic actuator.
Meissner and coworkers [77,80] have described a sliding plate
rheometer that can generate shearing in two perpendicular directions by use of two electromechanical linear actuators. The two
relevant components of the shear stress are measured independently using a shear stress transducer of the type described in the
next section.
It does not appear to be possible to make a direct measurement
of normal stress differences using the sliding plate geometry. A
pressure transducer can be mounted in one plate to determine (7"22'

292

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

but all and a 33 are unknown because of the nonhomogeneous flow


and stress fields near the free surfaces of the sample. However,
sliding plate rheometers have been used to measure birefringence
in bulk polymers [76] and in polymer solutions [71,87]. In this way,
information about normal stress differences can be obtained by use
of the stress optical law [88]. By the use of glass windows or entire
plates made of glass, the measurement of birefringence in the
1,3-plane is straightforward [87].
7.9.4 The Shear Stress Transducer

Many of the possible sources of error associated with the use of


sliding plate rheometers can be eliminated or dramatically reduced
if the shear stress is measured locally, at the center of the sample,
rather than being inferred from the total driving force, F. This is
especially advantageous in the study of the response of viscoelastic
materials to large strains.
First, such a technique is immune to the effects of instrument
friction. Thus, the support/guide mechanism for the moving plate
can be adjusted to have minimum play without regard to the
introduction of significant sliding friction. It is thus unnecessary to
use a "sandwich" configuration, and sample preparation and insertion are much simplified. Knowledge of the exact wetted area is no
longer required, as only the area of the shear sensitive surface of
the transducer needs to be known to calculate the shear stress.
A related advantage is that changes in the sample occurring at its
free surfaces due to oxidation and loss or absorption of water or
solvent have little effect on the stress and strain in the neighborhood of the shear stress transducer for an extended period of time.
By contrast, in cone-plate and parallel disk rheometers, the free
surface at which degradation and composition changes first occur is
at the outside radius where it has the maximum effect on the
measured torque and normal force. Another important advantage
of local shear stress measurement, especially at very high strains, is
the elimination of end effects so that the true shear stress response
to very large deformations can be determined.
A shear stress transducer suitable for use in the study of the
viscoelastic behavior of highly viscous liquids has been developed by
Dealy [89]. Giacomin and Dealy [69,90,91] used a transducer of

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

293

this type in a sliding plate rheometer designed for the study of


nonlinear viscoelasticity in molten plastics. A precision linear bearing table is used to support the moving plate. Samples are prepared
in the form of rectangular plaques about 1 mm thick, and only a few
grams are required. The rheometer is easily opened for cleaning
and sample insertion. Studies can be carried out at temperatures up
to 350C.
The transducer is entirely within the oven and is thus at a known,
uniform temperature. It can be calibrated at operating temperature
while mounted in the rheometer. The compliance of the rheometer
frame is negligible, and the actuator displacement is thus proportional to the sample deformation at all times. However, the shear
stress sensitive element of the transducer is compliant and has a
response time that depends on its stiffness and shape and on the
rheological properties of the melt. The time constant can be minimized by use of an elastomeric seal, or the compliance can be
minimized by use of a servo-mechanism [89]. However, the combination of a stiff transducer and a capacitance proximity probe yields
a satisfactory transducer response without these design complications. If the transducer is not sealed, it must be cleaned at regular
intervals to remove accumulated polymer.
The moving plate is driven by a servohydraulic linear actuator
under computer control, so that any type of strain history can be
programmed. This drive system can generate total strains up to 400
and shear rates up to 200 s - 1. The rheometer can be used to
measure the complex modulus over several decades of frequency
and the viscosity over several decades of shear rate. In addition a
wide range of nonlinear viscoelastic properties can be determined.
Examples of deformations useful in this regard are multiple step
strains, start-up and cessation of steady shear, interrupted shear,
large amplitude oscillatory shear, and exponential shear. Giacomin
et al. [92] have presented results of such tests for several polyethylenes.
Meissner et al. [80] have used a shear stress transducer in a
bidirectional shearing apparatus to measure simultaneously the
shear stresses in two directions.
For studies involving very high shear rates, sliding plate rheometers are at a disadvantage, because the total strain is limited by the
maximum displacement of the movable plate. In the "Sliding Film

294

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Rheometer" this problem is overcome by the use of a sandwich


geometry with a very long, flexible steel tape acting as the central
plate [93]. The tape moves at high speed from a feed reel to a
motor-driven take-up reel. The reservoir is continuously supplied
with melt by an extruder, and the shear stress is inferred from the
shear force on one of the side walls.
7.10 SLIDING CYLINDER RHEOMETERS

Sliding cylinder rheometers have no edge effects and require no


bearings or bushings to maintain the gap spacing. They have been
used to study molten plastics [94, 95] and polymer solutions [96].
The basic equations and sources of error for this geometry have
been presented in other books [1,2]. Because of sample loading and
cleaning problems, this type of rheometer has not been widely used
with molten polymers.
McCarthy [97] studied the dynamic mechanical properties of
molten polymers by forcing the inner cylinder to oscillate in the
axial direction. His rheometer incorporated a novel sample loading
technique that allowed the direct use of resin pellets. The oscillation of the inner cylinder was driven by an MTS servohydraulic
linear actuator.
Maxwell [95, 98] has developed a very simple sliding cylinder
rheometer for the study of stress relaxation after cessation of steady
shear in molten plastics. In this rheometer, the outer cylinder is
driven at constant speed, while the central rod is held in place by a
force transducer.
REFERENCES

1. J. M. Dealy, Rheometers for Molten Plastics, Van Nostrand Reinhold, N.Y.


(1982).
2. A. A. Collyer and D. W. Clegg, editors, Rheological Measurement, Elsevier
Applied Science, London & New York, 1988.
3. M. A. Lockyer and K. Waiters, Rheol. Acta 15:179 (1976).
4. H. M. Laun, 1. Rhea!. 30:459 (1986).
5. J. Meissner, 1. App!. Polym. Sci. 16:2877 (1972).
6. D. Chan and R. L. Powell, 1. Rheol. 28:449 (1984).
7. K. H. Lee, L. G. Jones, K. Pandalai and R. S. Brodkey, Trans. Soc. Rheol.
14:555 (1970).

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

295

8. G. K. Batchelor, An Introduction to Fluid Mechanics, p. 191, Cambridge Univ.


Press, Cambridge (1967).
9. Y. Mochimaru, J. Non-Newt. Fl. Mech. 12:135 (1983).
10. A. Narain and D. D. Joseph, Rheol. Acta 21:228 (1982).
11. A. W. Chow and G. G. Fuller, J. Non-Newt. Fl. Mech. 17:233 (1983).
12. I. F. MacDonald, B. C. Marsh and E. Ashare, Chem. Eng. Sci. 24:1615 (1969).
13. J. L. Schrag, Trans. Soc. Rheol. 21:399 (1977).
14. G. Heuser and E. Krause, Rheol. Acta 18:553 (1979).
15. J. M. H. Fortuin, Chem. Eng. Sci. 40:111 (1985).
16. P. T. Gavin and R. W. Whorlow, J. Appl. Polym. Sci. 19:567 (1975).
17. D. S. Pearson and W. E. Rochefort, J. Polym. Sci. Polym. Phys. 20:83 (1982).
18. L. M. Quinzani and E. M. Valles, J. Rheol. 30:S1 (1986).
19. W. Gleissle, Colloid & Polym. Sci. 252:848 (1974).
20. W. Gleissle, Rheol. Acta 15:305 (1976).
21. F. Nazem and M. G. Hansen, J. Appl. Polym. Sci. 20:1355 (1976).
22. J. F. Hutton, Rheol. Acta 8:54 (1969).
23. R. I. Tanner and M. Keentok, J. Rheol. 27:47 (1983).
24. C. W. Macosko and D. J. Morse, Proc. VIlth Intern. Congr. Rheol., p. 376
(1976).
25. A. S. Lodge, J. Non-Newt. Fl. Mech. 14:67 (1984).
26. I. Bruker, Rheol. Acta 25:501 (1986).
27. N. Adams and A. S. Lodge, Phil. Trans. A256:149 (1964).
28. W. Gleissle, Proc. VIIth Intern. Congr. Rheol., Gothenberg, Sweden, p. 594,
1976.
29. R. L. Crawley and W. E. Graessley, Trans. Soc. Rheol. 21:19 (1977).
30. L. J. Zapas, G. B. McKenna and A. Brenna, J. Rheol. 33:69 (1989).
31. I. Bruker and A. S. Lodge, J. Rheol. 29:557 (1985).
32. J. Meissner, R. W. Garbella and J. Hostettler, J. Rheol. 33:843 (1989).
33. K. Walters and R. A. Kemp, Rheol. Acta 7:1 (1968).
34. M. M. Cross and A. Kaye, Polymer 28:435 (1987).
35. K. Geiger, Rheol. Acta 27:209 (1988).
36. E. Ganani and R. L. Powell, J. Rheol. 29:931 (1985).
37. P. R. Soskey and H. H. Winter, J. Rheol. 28:625 (1984).
38. W. C. MacSporran and R. P. Spiers, Rheol. Acta 21:193 (1982).
39. W. C. MacSporran and R. P. Spiers, Rheol. Acta 23:90 (1984).
40. R. L. Powell and W. H. Schwartz, J. Polym. Sci. Polym. Phys. 17:969 (1979).
41. R. L. Powell and W. H. Schwartz, J. Rheol. 23:323 (1979).
42. B. Maxwell, Polym. Eng. Sci. 7:145 (1967).
43. R. J. J. Jongschaap, K. M. Knapper and J. S. Lopulissa, Polym. Eng. Sci.
18:788 (1978).
44. T. N. G. Abbott and K. Walters, J. Fl. Mech. 40:205 (1970).
45. P. Payvar and R. I. Tanner, Trans. Soc. Rheol. 17:449 (1973).
46. M. Gottlieb and C. W. Macosko, Rheol. Acta 21:90 (1982).
47. C. Goldstein and W. B. Schowalter, Trans. Soc. Rheol. 19:1 (1975).
48. L. H. Gross and B. Maxwell, Trans. Soc. Rheol. 16:577 (1972).
49. J. S. Dodge and I. M. Krieger, Rheol. Acta 8:480 (1969).

296

50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.
78.
79.
80.
81.
82.
83.
84.
85.
86.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

T. M. T. Yang and I. M. Krieger, J. Rheol. 22:413 (1978).


T. T. Tee and J. M. Dealy, Trans. Soc. Rheol. 19:595 (1975).
S. Onogi, T. Masuda and T. Matsumoto, Trans. Soc. Rheol. 14:275 (1970).
T. Matsumoto, Y. Segawa, Y. Waroshina and S. Onogi, Trans. Soc. Rheol.
17:47 (1973).
S. Onogi and T. Matsumoto, Polyrn. Eng. Rev. 1:45 (1981).
J. M. Dealy and T. K. P. Vu, J. Non-Newt. Fl. Mech. 3:127 (1977/78).
B. Maxwell, Plastics Engineering, Sept. 1987, p. 41.
D. J. Plazek, J. Polyrn. Sci. A-2 6:621 (1968).
G. Link and F. R. Schwarzl, Rheol. Acta 24:211 (1985).
T. E. R. Jones, J. M. Davies and A. Thomas, Rheo/. Acta 26:14 (1987).
A. J. P. Frank, J. Rheol. 29:833 (1985).
R. F. Garritano, U.S. Patent 4,501,155 (1985).
H. M. Laun and J. Meissner, Rheo/. Acta. 19:60 (1980).
W. Philippoff, in Physical Acoustics, II-B, Ed. W. P. Mason, Academic Press,
N.Y. (1965).
W. T. Read, J. App/. Mech. 17:349 (1950).
R. S. Rivlin and D. W. Saunders, I. R. I. Trans. 24:296 (1949).
E. N. da C. Andrade, Proc. Roy. Soc. Lond. 85:448 (1911).
A. N. Gent, R. L. Henry and M. L. Roxbury, J. App/. Mech. 41:855 (1974).
S. Toki and J. L. White, J. Appl. Polyrn. Sci. 27:3171 (1982).
A. J. Giacomin, Ph. D. Thesis, Chern. Eng., McGill Univ., Montreal, 1987.
A. J. Giacomin and J. M. Dealy, "The effect of free boundaries in melt
rheometers", Paper G3, 58th Annual Mtg., Soc. Rheo!., Tulsa, Oct. 1986.
K. Osaki, S. Kimura and M. Kurata, J. Rheol. 25:549 (1981).
T. Y. Liu, D. W. Mead, D. S. Soong and M. C. Williams, Rheo/. Acta 22:81
(1983).
N. Sivashinsky, A. T. Tsai, T. J. Moon and D. S. Soong, J. Rheol. 28:287
(1984).
T. Y. Liu, D. S. Soong and M. C. Williams, J. Polyrn. Soc. Polyrn. Phys.
22:1561 (1984).
J. M. Dealy and S. S. Soong, J. Rheol. 28:355 (1984).
S. K. Kimura, K. Osaki and M. Kurata, J. Polyrn. Sci. Polyrn. Phys. 19:151
(1981).
J. Meissner, Chernie 38:35 (1984).
F. P. LaMantia, B. de Cindio, E. Sort a and D. Acierno, Rheol. Acta 18:369
(1979).
D. Acierno et a!., Trans. Soc. Rheo/. 21:261 (1977).
J. Meissner, B. Ziille and H. Hiirlimann, Proc. Xth Intern. Congr. Rheo!. 2:121
(1988).
H. M. Laun, Rheol. Acta 21:464 (1982).
J. Meissner, Pure App/. Chern. 56:369 (1984).
M. Fleissner, Angewandte Makrorn. Chernie. 94:197 (1981).
F. Nazem and Y. Hill, Trans. Soc. Rheol. 18:87 (1974).
B. H. Shah and R. Darby, Polyrn. Eng. Sci. 16:46 (1976).
T. Murayama, J. App/. Polyrn. Sci. 27:80 (1982).

ROTATIONAL AND SLIDING SURFACE RHEOMETERS

297

87. A. Haghtalab, M. Eng. Thesis, Chern. Eng., McGill Univ., Montreal, 1985.
88. J. Janeschitz-Kriegl, Polymer Melt Rheology and Flow Birefringence, Springer
Verlag, Berlin, N.Y. (1983).
89. J. M. Dealy, U.S. Patent No. 4,463,928 (1984).
90. A. J. Giacomin and J. M. Dealy, SPE (ANTEC) Tech. Papers 32:711 (1986).
91. J. M. Dealy, U.S. Patent No. 4,571,989 (1986).
92. A. J. Giacomin, T. Samurkas and J. M. Dealy, Polym. Eng. Sci. 29:499 (1989).
93. F. C. Starr, U.S. Patent No. 4,466,274 (1984).
94. A. W. Myers and J. A. Faucher, Trans. Soc. Rheol. 12:183 (1968).
95. B. Maxwell, in Order in the Amorphous State of Polymers, ed. by S. E. Keinath,
R. L. Miller and J. K. Rieke, Plenum Publ. Corp., N.Y. (1986).
96. A. T. Tsai and D. S. Soong, J. Rheol. 29:1 (1985).
97. R. V. McCarthy, J. Rheo/. 22:623 (1978).
98. B. Maxwell and K. S. Cook, J. Polym. Sci. Polym. Symp. 72:343 (1985).

Chapter 8
Flow in Capillaries, Slits
and Dies
8.1 INTRODUCTION

Pressure driven flow through tubes, slits and other types of channels is of central importance in experimental rheology and in
polymer processing. Not only is this flow used as the basis for the
most popular type of melt rheometer, but it is also a flow that
occurs often in melt processing, for example in an extrusion die or
in the runner feeding an injection mold. We will derive the basic
equations for flow in tubes and slits and show how these can be
used to interpret rheometer data and to design flow systems. The
irregular flows that can occur at the entrance and exit of a die are
described, and methods for estimating the pressure drop in dies are
reviewed.
8.2 FLOW IN A ROUND TUBE
8.2.1 Shear Stress Distribution

For the usual cylindrical coordinate system, there is only one


independent nonzero shear component of the extra stress tensor,
and we will use the symbol, 0-, for this. Taking into account the
symmetry of the stress tensor, we have:
(8-1 )
For the steady flow of an incompressible fluid in a tube of radius,
R, driven by a pressure gradient (dP /dz), a force balance on a
298

FLOW IN CAPILLARIES, SLITS AND DIES

299

cylindrical element of fluid gives:

a(r)
At the wall, r

"2r (dP)
dz

(8-2)

R(dP)
-

(8-3)

R, and:

a(R) = -

2 dz

Thus:
r

(8-4)

a(r) = R a(R)

Note that this analysis is independent of the rheological properties


of the fluid, so that the resulting equations are valid for both
Newtonian and non-Newtonian fluids.
jdz) is negative (the pressure "drops" in the flow
Since
direction), a(R) is also negative. For convenience, we define a
"wall shear stress," which has a positive value:

(dP

a: = -a(R) = R (_

dP)
dz

(8-5)

8.2.2 Shear Rate for a Newtonian Fluid

For fully developed flow, i.e., flow in which all streamlines are
parallel to the tube axis, the shear rate is:

Y == Yrz

TJ'Y

dv
dr

(8-6)

For a Newtonian fluid:


=

(8-7)

where TJ is a constant. Combining Equation 8-7 with Equation 8-2

300

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

we have:

y(r) = - r

21'/

(dP)
dz

r
= -y(R)
R

(8-8)

Combining Equations 8-6 and 8-8, and assuming that the velocity at
the wall is zero (no slip), we can obtain the velocity profile:
(8-9)
where Q is the volumetric flow rate. The shear rate at the wall for a
Newtonian fluid is:

y(R) =

(ddr
V

(8-10)

r=R

The shear rate at the wall is negative. For convenience, we represent the magnitude of this quantity by the positive quantity, Yw:

(8-11)
Thus,
yw(Newtonian fluid) = 4Q3
nR

(8-12)

For non-Newtonian fluids, the velocity profile is no longer given


by Equation 8-9, and the shear rate at the wall is not given by
Equation 8-12. However, the quantity on the right in the latter
equation is still useful, and we define it as the "apparent shear rate
at the wall," YA'

(8-13)
Thus, for a Newtonian fluid, a plot of U w (as given by Equation 8-5)
versus YA will be a straight line with a slope equal to the viscosity.
On a log-log plot, the slope will be one.

FLOW IN CAPILLARIES. SLITS AND DIES

301

Dividing Equation 8-5 by Equation 8-12, we obtain an equation


for calculating the viscosity of a Newtonian fluid from tube flow
data:
7](Newtonian fluid)

( -dpjdz )7TR 4
8Q

UW

= -.

'Yw

(8-14)

For a non-Newtonian fluid, we call this quantity the "apparent


viscosity":
( -dpjdz)7TR4

8Q

(8-15)

8.2.3 Shear Rate for a Power Law Fluid

For a non-Newtonian fluid the shear stress no longer varies linearly


with shear rate, and the viscosity as a function of the shear rate
must be known to derive an expression for y(r). The power-law
model has proven useful for calculating the velocity and shear rate
distribution for tube flow of shear thinning fluids. While this model
is incorrect at low shear rates, in tube flow this only affects the
predicted velocity near the axis, and the resulting error in the
velocity profile is quite small.
The shear stress is related to the shear rate as follows according
to the power law:

(4-9)
To avoid problems with nonintegral powers of negative numbers,
we will write this as follows for tube flow:

-u=K(-yr

(8-16)

Making use of Equations 8-2, 8-6 and 8-16 we can obtain a


relationship between the pressure gradient and the velocity
gradient:
(8-17)

302

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

3.5 ...---,-"""T"-""T""-"T"'""--r--r-----,-.....,..--r---,
3.0 ~---l-_+--+-+-____:!--r-----f--t---+---t

oi= 2 .0 1-----'1----1---+-#+~-+-.....:;::r-~+_-+_-t-__1
<{

a:

~ 1 . 5~~~-+~~--_+_-~-~--~~+_-t__;

~ 1 . 0 ~~~~-_+-_+---+~~-~-~~+__;

>

0.5 hf.#J.f--f--+--I---+---+---+--+--ft:\--l

0.8

0.6
RADIUS RATIO. rl R

Figure 8-1. Tube flow velocity profiles calculated from the power law for several values of n.
Plot shows v(r) IV as a function of r I R ,where V is the mean velocity.

Again taking the velocity at the wall to be zero (no slip), the velocity
profile can be determined by integrating Equation 8-17.
Profiles for several values of n are shown in Figure 8-1. Note that
for values of n typical of many high molecular weight polymers, the
shear rate is high only in a very narrow region near the wall and is
relatively low over the central portion of the tube. Furthermore,
since the velocity is low near the wall, where the shear rate is
highest, during any interval of time, the melt leaving the die will
contain relatively little material that has been sheared at a rate
close to YW. This means that the total strain that has been experienced by the melt as it exits the tube is highly nonuniform, and that
while the wall shear rate is the maximum value, it is far from the
average value for the melt leaving the tube.
The shear rate at the wall is given by:

. = (3n + 1 ) ( ~) = ( 3n + 1 ) .
4n
'TT"R 3
4n
'YA

'Yw

(8-18)

FLOW IN CAPILLARIES. SLITS AND DIES

303

Thus, the quantity (4Q/,rrR 3 ), which is the wall shear rate for a
Newtonian fluid, no longer has this significance.
Equation 8-18 can be used to calculate the error involved in using
the apparent wall shear rate as an estimate of the true value. For
example, when n = 0.5, the actual wall shear rate is 1.25YA'
Combining Equation 8-18 with Equation 4-9 we obtain:
O"W =

K(

3n +
4n

1)n . n
('YA)

(8-19)

Thus, a plot of log o"w versus log YA is a straight line with a slope
of n.
8.2.4 The Rabinowitch Correction

If a specific form of the relationship between the shear stress and

the shear rate is not assumed, then it is not possible to calculate the
true shear rate at the wall directly, knowing only YA' However,
there is a way to determine values of the true wall shear rate and
the viscosity, even when no viscosity function is assumed, providing
that the pressure drop has been determined at a number of flow
rates. The equations required to carry out this calculation are
derived in many books, for example those of Walters [1] and
Whorlow [2], and we will only summarize the results here.
First, it can be shown that for a specific fluid and temperature,
there is a unique relationship between the shear stress at the wall
and the apparent wall shear rate. This means that if we obtain
pressure drop data for a variety of flow rates, tube lengths and tube
radii, they will all fall on a single curve when a plot of [4Q/7TR 3 ]
versus [( - LlP)R/2L] is prepared. A double logarithmic plot is
often used, and it can be further shown that the shear rate at the
wall is related to the slope of the curve on such a plot.

Yw =

(3 : b )YA

(8-20a)

where:
(8-20b)

304

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

This relationship has been variously associated with the names of


Weissenberg, Rabinowitch, and Mooney, while the bracketed term
in Equation 8-20a is usually called the Rabinowitch correction. It is
generally found to be independent of temperature. By comparison
with Equation 8-12, we see that this term represents the deviation
from Newtonian behavior. For a power law fluid, it is obvious from
Equation 8-18 that b = lin.
From the above discussion we see that a plot of the logarithm of
the apparent wall shear rate versus the logarithm of the wall shear
stress reveals the general nature of the viscosity function. If the
data fall on a straight line with a slope of 1, Newtonian behavior is
indicated. If they fall on a line with a slope not equal to 1, then
power law behavior is indicated, and the slope is equal to lin.
Curvature implies that the behavior is neither Newtonian nor
power law, and Equation 8-20 must be used to determine the true
wall shear rate.
Once this has been done the viscosity can be calculated. Thus,
the above equations provide a basis for the determination of the
viscosity of a non-Newtonian fluid on the basis of tube flow data.
However, this requires differentiation of the data so that it is not
possible to calculate a true value of the viscosity using data from a
single experiment.
8.2.5 The SchOmmer Approximation

Schlimmer et al. [3,4] have proposed an approximate procedure for


determining the viscosity function from tube flow data without
differentiation of the data and determination of the Rabinowitch
correction. They define a "representative shear rate," y*, as some
fraction, x*, of the apparent wall shear rate:
(8-21)
This is the shear rate at which, for a given flow rate, the true
viscosity would be equal to the apparent viscosity:
(8-22)

FLOW IN CAPILLARIES. SLITS AND DIES

305

For a power-law fluid, it can be shown that, for capillary flow,


x* = (

3n + l)n/(n-t>
4n

(8-23)

For values of n in the range of 0.36 to 1.2, x* is a weak function


of n:
x* = 0.83

4%

(0.36 < n < 1.2)

And the resulting error in the viscosity is less than 3%. Schiimmer
proposes that this constant value be used, thus making it possible to
calculate the viscosity at one shear rate using data from a single
experiment. Or, the entire curve of TJA versus YA on a log-log plot
can be shifted horizontally to obtain the curve of TJ( Y)'
8.2.6 Wall Slip in Capillary Flow

At a critical value of the wall shear stress, it has been observed that
the melt no longer adheres to the wall, and we present here the
equations used to describe wall slip in a capillary. We allow the
velocity of the melt at the wall to be V:, rather than zero, as is
assumed in classical fluid mechanics. Thus, the equations involving
the velocity given above are modified by replacing v by (v - v:),
which is the melt velocity relative to its velocity at the wall. In
particular, the equation for the wall shear rate must be altered.
First, we note that the volumetric flow rate, Q, can be written in
terms of the average velocity, V, as follows:
(8-24)

Thus, the apparent wall shear rate, YA' as defined by Equation 8-13,
is:

.
'V

where D

2R.

4Q

4V

8V

'lTR 3

R -

=--=---

fA -

(8-25)

306

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

When wall slip occurs, V in this equation must be replaced by


(V - ~) to obtain an apparent wall shear rate valid when slip is
occurring, which we shall call YA,S.
'YA, S

8(V - ~)

(8-26)

Thus, according to the argument presented in Section 8.2.4, for a


given resin and temperature, there is a unique relationship between
U w and YA,S. However, if slip is not accounted for, and U w is plotted
as a function of YA for capillaries of various diameters, the data will
not collapse onto a single curve. Such an observation would suggest
the occurrence of slip. If the values of YA obtained with various
capillary diameters, for a fixed value of U w (and thus a fixed value of
YA,S)' are then plotted versus 11R (or liD), a straight line will
result, as indicated by a rearrangement of Equation 8-26:
.

'YA

= 'YA,S +

4~

(8-27)

Ramamurthy [5] has shown a set of such lines for an LLDPE for
different values of wall shear stress. The slope of each of these lines
is 4~, and this relationship thus provides a basis for the experimental determination of slip velocity as a function of wall shear stress.
For a power-law fluid, YA can be replaced by YA,S in Equation
8-19 to give:
Uw =

K(

3n+1)n(.
4n
'YA

4~)n
R

(8-28)

This can be solved for the slip velocity to give:

4~
R

~
7TR 3

(Uw )lln(
K

4n )
3n + 1

(8-29)

Thus, if the power law constants have been determined in experiments in which slip did not occur, they can be used to calculate the
slip velocity from a measured value of U w for any given value of Q.

FLOW IN CAPILLARIES, SLITS AND DIES

307

Values of the slip velocity determined in this way have been


reported for LLDPE by Kalika and Denn [6].
If log(lTw) is plotted versus loge YA)' the data will start to fall below
the power law line when slip occurs, as indicated by Equation 8-28.
At the same time, the surface of the extrudate begins to be rough, a
phenomenon discussed in detail in Section 8.9. Ramamurthy [5] has
discussed the occurrence of slip in a capillary and its effect on the
lTw( YA) curve and the extrudate appearance.
The effect of slip on the flow curve is not a dramatic one, and it is
generally not possible to detect slip solely by examination of this
curve. The use of capillaries of several radii together with Equation
8-27 is a much more conclusive method for demonstrating that slip
has occurred.
8.3 FLOW IN A SLIT

A slit is defined here as a straight, rectangular channel having a


width, W, that is much greater than its thickness, h. For such a flow
geometry, the edges make a negligible contribution to the pressure
drop, and the fully developed flow can be considered to be twodimensional. The flow field in certain commercial forming operations approximates slit flow, and this geometry is also used for
rheological measurements. In this regard, it has certain advantages
over capillary flow. First, flush-mounted pressure transducers can
be used. This eliminates the need for end corrections. Second, the
two-dimensional flow field facilitates the observation of flow in the
channel and at its entrance and exit and the use of optical techniques such as birefringence.
8.3.1 Basic Equations for Shear Stress and Shear Rate

A force balance on a rectangular element of fluid in the slit yields a


relationship between the shear stress and the pressure gradient:

IT(Y) == TyxCy)

!::.p .y
=

(8-30)

where y is measured from the center plane of the slit as shown in

308

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

~---------L----------~

Figure 8-2. Slit flow geometry showing the meaning of the symbols used in the text.

Figure 8-2. The magnitude of the shear stress at the wall,


then:
O"w =

-O"(y =

h/2)

(--IlP)h
L - "2

O"w,

is

(8-31)

The above equations for the shear stress are valid for any fluid,
but in order to relate the shear rate at the wall to the flow rate, the
viscosity must be known as a function of the shear rate. For
example, for a Newtonian fluid, the magnitude of the shear rate at
the wall is:
yw(Newtonian fluid)

-YyAy

h/2)

6Q/h 2 w (8-32)

and the viscosity can be calculated as follows:


'YI

O"w _

--

. , .Y

-IlP)
hw
L
12Q
3

-----

(8-33)

If the fluid is non-Newtonian, the shear rate at the wall is no


longer given by Equation 8-32. In this case, it is not possible to
calculate Yw directly from Q. However, if the pressure drop has
been measured as a function of flow rate, the wall shear rate can be

FLOW IN CAPILLARIES, SLITS AND DIES

309

determined by means of Equation 8-34, a derivation of which is


given by Walters [1]:
(8-34a)
where:
(8-34b)
As in the case of capillary rheometers, a plot of the logarithm of the
apparent wall shear rate (6Q/h 2 w) versus the logarithm of the wall
shear stress, a w , reveals the general nature of the behavior of the
function. If the data fall on a straight line with a slope of 1,
Newtonian behavior is indicated. If they fall on a straight line
with a slope not equal to 1, then power law behavior is indicated,
and n = 1/f3. Curvature indicates that the behavior is neither
Newtonian nor power law, and Equation 8-34 must be used to
determine Yw'
In Section 8.2.5, we described Schiimmer's procedure for obtaining a reasonable estimate of the viscosity curve using capillary flow
data, without differentiating the data. A similar procedure can be
used for slit flow data. In this case the quantity x* is given by:
x*

2n

+ 1 )n/(n-n

3n

::::;

0.79

(8-35)

Laun [7] has used this technique to determine the viscosity curve for
LDPE using slit data.
8.3.2 Use of a Slit Rheometer to Determine N,

Two methods have been proposed to determine the first normal


stress difference by measuring various pressures generated in slit
flow. One of these is the "hole pressure" method, and the other is
the "exit pressure" method. These techniques are described below.

310

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

8.3.2.1 Determination of N j from the Hole Pressure

It has been established that the pressure measured at the bottom of


a small transverse slot or circular hole in the wall of a flow channel
can be different from the pressure that would be exerted by the
fluid at the same point in the wall if the slot or hole were not
present. In the case of flow in a slit, this pressure difference is
practically equivalent to the difference between the pressure at the
bottom of a hole or slot and that indicated by a pressure transducer
mounted flush with the opposite wall of the slit. Lodge [8] calls this
difference the "hole pressure," which is defined as follows:
(8-36)
where:

PI
P2

=
=

pressure indicated by flush-mounted transducer


pressure at the bottom of the slot or hole

In particular, when the Reynolds number is quite low so that


inertial effects can be neglected, the hole pressure is found to be
zero for Newtonian fluids but positive for polymeric liquids.
Of interest here is the possible relationship between P* and N I .
Figure 8-3 illustrates a physical explanation of the existence of such
a relationship. In the neighborhood of the hole, the streamlines are
P,

1
STREAMLINE

7
_F~

~F

' ) ~L~~LS

~-- ~r------SLOT OR
HOLE

P2

Figure 8-3. Typical streamline near a hole or transverse slot and the forces that result when
Nl is positive.

FLOW IN CAPILlARIES. SLITS AND DIES

311

curved, and a positive NI implies the existence of forces in the


direction shown, which will tend to draw fluid out of the hole and
reduce P z to a value below Pl.
A detailed theoretical analysis of this flow requires the use of a
nonlinear constitutive equation and a numerical analysis. Such an
analysis does not lead to an analytical relationship between P* and
N I , but two approximate theoretical analyses have been used to
derive such relationships. The second order flow perturbation approach, valid for flows in which only a minute deviation from
inelastic, Newtonian behavior occurs, predicts [9], for a transverse
slot or a hole:
(8-37)
Higashitani and Pritchard [10] assumed that the flow near the hole
is approximately a viscometric flow and symmetric about the hole or
slot center. Their analysis gives the following results:
Transverse slot: P*

Circular hole: P* =

jU (NI /2u) du

(8-38)

jUW[(NI -Nz )/3u] du

(8-39)

These equations cannot be used to determine NI or N z directly


from measurements of P*, but they have been used to test the
theory by comparing measured values of P* with values calculated
from Equation 8-38 using NI(u) data obtained from a cone-plate
rheometer or by use of birefringence measurements across a slit on
the centerline of a slot [11, 12].
Baird [13] has shown that Equations 8-38 and 8-39 can be
differentiated to give:
Transverse slot:
Circular hole:

(NI

NI

2uw (dP* /duw )

(8-40)

Nz)

3uw (dP* /duw )

(8-41)

where the normal stress differences are the values corresponding to


the shear stress, U w

312

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Baird [14] further noted that if it is assumed that:

then:
Transverse slot: Nl

2mP*

(8-42)

Circular hole: Nl

3mP*

(8-43)

We note that these equations imply that:


d(log P*)
d(log O"w)

=m

(8-44)

The approximate analyses suggest that hole pressure data can be


used to calculate the normal stress differences. While the second
order analysis is very limited in its applicability, the Higashitani and
Pritchard theory is more general. However, the validity of its
underlying assumptions can only be established experimentally, and
careful flow birefringence studies have revealed that these assumptions are not valid [11]. Numerical simulations [15, 16] support this
observation. It is, therefore, surprising that the limited experimental data available for melts [12, 17] indicate a reasonable degree of
agreement between values of Nl calculated from hole pressure data
obtained using transverse slots and those determined by other
methods. Pike and Baird [11] found in their birefringence measurements that, at least for some melts, the departures from the
assumptions of the theory tend to cancel out in taking the integral
indicated by Equation 8-38. Whether or not this fortuitous situation
arises for a wide range of polymers remains to be learned.
Lodge [8] has reviewed in detail all the existing experimental
evidence regarding the determination of normal stress differences
by means of hole pressure measurements. While he finds cause for
some optimism regarding the existence of a simple, approximate
relationship between P* and Np the data currently available for
melts are insufficient to provide a solid foundation for the use of
this method.
Obviously this is an area that merits further research. For those
wishing to use the method, however, great care is necessary to

FLOW IN CAPILLARIES. SLITS AND DIES

313

assure an accurate determination of P*. Lodge [8] has examined


the possible sources of error in great detail. While inertial corrections are not important for melts, great care is required to detect
the difference between two relatively high pressures. Furthermore,
it is essential to avoid any leakage flow from the slot. Srinivasan and
Finlayson [18] have studied the effect of temperature non uniformity
and have provided corrections to P* to account for non-isothermal
conditions.
8.3.2.2 Determination of Nt from the Exit Pressure

If the flow in a slit is assumed to be fully-developed, viscometric


flow right up to the exit, it can be shown [19,20] that the wall
y), as follows:
pressure at the exit, Pe , is related to

Nt<

(8-45)
An equivalent representation is:

(8-46)

Based on this relationship, Han [21] has proposed the use of an


experimental technique for the determination of Nl in molten
polymers at high shear rates. Noting that the wall pressure cannot
be measured right at the exit, he suggested that wall pressures be
measured at several points along the flow direction and extrapolated to z = L to determine Pe , as shown in Figure 8-4. A commercial slit rheometer designed to make use of this technique is the
Seiscor /Han Rheometer made by Seiscor (see Appendix E).
There are two major sources of uncertainty in this procedure.
First, while the derivation of Equation 8-45 assumes that the
streamlines are straight and parallel right up to the exit, both
experimental studies [22,23] and numerical simulations [24, 25] show
that there is a velocity rearrangement that becomes significant some
distance upstream of the exit and that this effect becomes more
prominent as the flow rate is reduced.

314

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

o~

~,,-p

__________________
z

Figure 8-4. Extrapolation of measured wall pressures (PI' P 2 , P3 ) to z = L to determine the


exit pressure, Pe'

Estimates of the effect of velocity profile rearrangement on the


relationship between the exit pressure and the first normal stress
difference have been carried out for a second order slow flow
approximation [22] and for a convected Maxwell model [25]. However, both of these analyses involve fluids in which the viscosity is
independent of shear rate. Vlachopoulos and Mitsoulis [26] carried
out a finite element analysis of exit flow assuming a viscoelastic
model in which the viscosity decreased with shear rate in a realistic
way and N J was proportional to a power of (T. They calculated the
pressure distribution along a slit and found that the calculated
values of Nl determined from extrapolated exit pressures were
somewhat above the values calculated directly from the constitutive
equation. However, as the wall shear stress increased, the values
obtained by the two procedures became more nearly equal.
It would be desirable to be able to establish a general criterion to
be used to predict when this velocity rearrangement would lead to
significant errors in the use of Equations 8-45 and 8-46. Boger and
Denn [27] carried out a thorough analysis of the flow, without
making any assumptions about the rheological behavior of the fluid,

FLOW IN CAPILLARIES, SLITS AND DIES

315

and found that while the error will grow as the flow rate decreases,
there is no universal criterion for evaluating the relative importance
of the error.
Based on his experience with several polyethylene and polystyrene
melts, Han [28] has proposed that exit disturbances can be neglected when lTw is greater than 25 kPa. However, in establishing
this criterion, Han compared values of Nl calculated using Equation 8-45 with those extrapolated from low shear rate data obtained
from a cone-plate rheometer. Such an extrapolation is shown in
Figure 8-5. This extrapolation was necessary to make a comparison,
because cone-plate data are only obtainable at very low shear rates.
The second source of uncertainty in the exit pressure method is
the determination of Pe by means of an extrapolation procedure. If
the flow is visco metric, and the viscosity does not vary with downstream position, the plot of wall pressure versus downstream position will be a straight line, and the extrapolation is, in principle,
straightforward. In practice, however, there are several pitfalls.
First, the exit pressure is generally small compared to the measured
wall pressures. This means that very precise values of the pressure
must be obtained. Also, the pressure should be measured at sufficient points (at least three) to ensure that the measured values truly
fall on a straight line. Even small errors in these values can lead to
meaningless values of Pe [12,29]. A second problem is that curvature of the P(z) data can arise from the dependence of viscosity on
temperature and pressure. Viscous heating will always produce an
increase in temperature along the flow and thus a decrease in the
viscosity. Han [28] has pointed out that this effect can be reliably
calculated so that the importance of this source of error can be
easily determined.
The role of the pressure dependence of the viscosity has been
examined by Laun [7]. He found that this effect produced significant curvature in the P(z) curve for LDPE when the driving
pressure was above 100 bars. Once this curvature became significant, he found that it was impossible to make a reasonable estimate
of Pe , even if the pressure dependence of the viscosity was taken
into account in the calculation of Pe Han [28] warns that nonlinear
P(z) data should not be used to determine Pe On the other hand,
Tuna and Finlayson [30] found that a quadratic fit gave positive,

316

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


103~--------------------------------------,

CALCULATED FROM
EXIT PRESSURE

10

CONE-PLATE
DATA

1.0

(J

(kPa)

Figure 8-5. Comparison of NI values for a LDPE determined using the exit pressure method
with an extrapolation of low shear rate values measured using a cone-plate rheometer. NI is
plotted as a function of shear stress. Adapted from Ref. 28. Copyright 1988 by Elsevier
Applied Science Publishers Ltd. Reprinted by permission.

reasonable values for Pe and was consistent with a generally accepted equation describing the effect of pressure on viscosity.
In conclusion, there is some theoretical basis for the use of the
exit pressure method to determine Nt, but the extrapolation procedure requires data of exceptional precision. Furthermore, as the
wall shear stress is reduced, the error associated with velocity
rearrangement near the exit increases in a manner that depends
very much on the nature of the flowing fluid. Han [28] has reviewed
the existing data and recommends that for melts the effect of the

FLOW IN CAPILLARIES, SLITS AND DIES

317

exit disturbance can be neglected when the wall shear stress is


greater than 25 kPa. However, as this stress is increased further,
both viscous heating and the effect of viscosity on pressure will
begin to cause curvature in the plot of pressure versus length.
When any curvature is evident in the data, no extrapolation should
be attempted.
8.4 PRESSURE DROP IN IRREGULAR CROSS SECTIONS

The accurate calculation of the pressure drop for flow through a


straight channel having an irregular cross section, for example a
star or square, requires numerical solution of the equations of
motion. While only the viscosity function needs to be taken into
account as long as the flow is rectilinear, the calculation is not
trivial, especially for non-Newtonian fluids. For this reason, several
simplified methods have been proposed. Kozicki et al. [31] used two
geometric constants to relate the shear rate at the wall to the flow
rate, while Miller [32] used only one constant. Liu and Hong [33]
compared these approximation schemes and found that both become less reliable as the fluid behavior becomes more non-Newtonian.
8.5 ENTRANCE EFFECTS

The details of the flow at the entrance to a channel are important


for several reasons. When a capillary or a slit is used for rheological
measurements, a flat (a = 90) entrance like that shown in Figure
8-6 is usually used, although a = 45 is also used. There is a large
pressure drop associated with the flow in the entrance region, and
this must be taken into account if the reservoir pressure is the
quantity measured to determine the wall shear stress. Moreover, it
has been proposed that the excess pressure drop at the entrance to
a capillary is itself a useful quantity that can be used to characterize
polymers.
When melt is processed by extruding it from a die, the pressure
drop at the entrance of the die is important for design purposes.
Furthermore, the flow pattern at the entrance can have an important effect on the appearance of the extrudate as it leaves the die.

318

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

(a)

(b)

Figure 8-6. Flow patterns at the entrance to a capillary or slit with a flat entrance. 8-6a (left)
shows smooth flow with no vortices. 8-6b (right) shows corner vortices.

8.5.1 Experimental Observations

The types of behavior observed when melt flows from a reservoir to


a much smaller capillary or slit are sketched in Figure 8-6. For
Newtonian fluids at low Reynolds number, and for some melts such
as HDPE and polypropylene, all streamlines stretch in a regular
way from the reservoir into the capillary or slit, as shown in Figure
8-6a. For certain other polymers, however, such as LDPE and
polystyrene, there are prominent stationary vortices in the corner
regions, as shown in Figure 8-6b. These flow patterns have been
described in detail by White and Kondo [34]. As the flow rate is
increased, these vortices become unstable. In the case of flow into a
slit, first one and then the other vortex becomes larger, with
material from each vortex periodically "shooting" into the main
flow. In the case of capillary flow, the instability leads to a timeunsteady, spiral flow or to a random "shooting" of vortex fluid into
the capillary. In all cases, the unstable flow at the entrance is
reflected in a marked distortion of the melt extruded from the
capillary or slit [35, 36].
White and Baird [37] found that for polystyrene flowing into a
slit, there is a moderate level of vortex flow at low temperatures
and small contraction ratios (viz. 4: 1), but these vortices don't grow
as the flow rate increases. White et al. [38] hypothesized that the
vortex type flow pattern occurs when the ratio of extensional to
shear stress is high.

FLOW IN CAPILLARIES, SLITS AND DIES

319

The slit flow birefringence studies of Han and Drexler [39] and
Aldhouse et al. [40] have shown that while the detailed stress
distribution at the entrance varies from one material to another, the
wall shear stress reaches its steady state value very quickly, usually
within a distance of O.th from the entrance, where h is the slit gap,
as shown in Figure 8-2. The tensile stress on the center plane
decays from a maximum value at the entrance plane to zero over a
distance that increases with flow rate. This distance is often as
much as several multiples of h, and one might presume that it is
proportional to AQ/A, for a given flow geometry, where A is a
relaxation time of the fluid, and A is the cross sectional area of the
slit.
Laser velocimetry measurements [4t] have shown that the velocity profile approaches its fully developed form very close to the
entrance, i.e., within a distance of about O.th. These general
observations regarding the stress and velocity distributions near the
entrance to a slit are thought to be valid also for capillaries.
Piau et al. [35] have made careful observations of entrance flows
of linear and branched silicone fluids (PDMS). They describe in
detail the various types of vortex structures that occur and relate
these to the rheological properties of the fluids.
8.5.2 Entrance Pressure Drop: The Bagley End Correction

Wall pressures measured at various axial locations in a reservoir


and capillary have been :reported by Han [42] for molten polymers.
A typical result is shown in Figure 8-7. We note that there is a large
pressure drop associated with the entrance region, but that the
pressure gradient in the capillary rapidly approaches a constant
value, usually within a dimensionless distance, z/R, of one. There
also appears to be a small residual pressure at the exit of the
capillary, which we have called the exit pressure.
The total pressure drop for flow from a reservoir, through a
capillary and out to the ambient pressure can be thought to consist
of three components:

(8-47)

320

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

700
0)600
iii
a.

~500

::::l

400

en

~ 300
II:

a.

:::l

,
,

-----<:>0,

\
--.&&., \

----G...

\
\ \
\ \

\.
..... \

----e.. \

"\\

200

> 100
o~~~~~~~~~~~~~~~~

- 0.4 - 0.3 - 0.2 - 0.1 0.0 0.1 0.2 0.3 0.4 0.5
DISTANCE FROM ENTRANCE (in.)

0.6

Figure 8-7. Pressure distribution in a reselVoir and capillary for HDPE. L = 0.5 in. and
D = 0.125 in.; Shear rates, in S-1 are, from top to bottom: 790, 616, 313 and 160. Adapted
from Ref. 71. Copyright 1970 by The Society of Rheology. Reprinted by permission of
John Wiley & Sons, Inc.

where:

Pd
Pa

ilPent
ilPcap

ilPex

=
=

driving pressure
ambient pressure to which capillary exhausts
excess pressure drop due to entrance flow
pressure drop for fully developed flow in a capillary
of length L
excess pressure drop due to exit flow

Because for Newtonian fluids at low Reynolds numbers, ilPent is


very small and ilPex is zero, the two "excess" pressure drops are
thought to reflect the elasticity of the melt.
When using capillary flow data along with the equations of
Section 8.2 to determine melt viscosity, one can determine the true
pressure gradient for fully developed flow, i.e., dP /dz in Equation
8-5, by measuring the wall pressure at various axial distances and
making a plot like that shown in Figure 8-7. However, it is preferable to be able to infer the viscosity from the simpler procedure of
measuring the driving pressure for various flow rates. Bagley [43]
has suggested a scheme for doing this. He measured the driving
pressure (Pd ) at various values of the flow rate (Q) using a variety
of capillaries having different lengths. For each value of the appar-

FLOW IN CAPILLARIES, SLITS AND DIES

o~

__

____

__

______________________

321

L/R

Figure 8-8. Bagley plot for determining the end correction for capillary flow. Data for two
values of YA are shown.

ent wall shear rate (4Qj7TR 3 ) he then plotted driving pressure


versus LjR and drew a straight line through the points as shown in
Figure 8-8- Extrapolating the lines corresponding to various values
of YA to the Pd = 0 axis, he then obtained an "end correction,"
"e," defined as the negative of the value of LjR at the point of
interception. This represents the length of the capillary (divided by
R) for which fully developed flow would give a pressure drop equal
to dPends ' Thus, the true wall shear stress can then be calculated as
follows:
Uw =

Pd
-2(-:-L-j-R-+-e-:-)

(8-48)

It can be shown that there is a simple relationship between the

Bagley end correction, e, and dPends :


(8-49)
For a Newtonian fluid, the equations of motion have been solved
analytically [44] to show that:
dPends

dPent

2.30uw (Newtonian fluid)

(8-50)

322

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

where (Tw is the true wall shear stress for fully developed tube flow.
This implies that e = 1.15.
Bagley found that his Pd versus L / R curves continued to be
linear even for very short dies. This is in accord with the observations reported above that the wall shear stress approaches its fully
developed flow value immediately downstream of the entrance. It is
interesting to note in this regard that for the flow of a Newtonian
fluid through a thin orifice [45],
IlPent

3YJQ/R3 (Newtonian fluid)

(8-51)

This implies that e = 1.18, which is close to the value of 1.15 cited
above for the Newtonian entrance correction for capillary flow.
Assuming that the similarity between orifice flow and capillary
entrance flow is also valid for melts, Cogswell [45] has recommended the use of one long die (L/R = 32) and an orifice (L/R z
0) as a shortcut method for determining reasonably accurate values
for (Tw. Thus, the driving pressure for the orifice plate, PJ, is taken
to be IlPends ' Laun and Schuch [46] found that the use of an orifice
(L/R < 0.2; entrance angle = 120) gave a IlP that agreed well
with IlPends determined by extrapolation of data from flat entry
capillary dies. Neglecting the exit pressure loss, the true wall shear
stress is then:

(T

(pJ - PJ)R
2L

(8-52)

and the end correction is:

PJL
(pJ - PJ)R

e = --,-----:---,----

(8-53)

In our discussion of end effects, the dependence of viscosity on


temperature and pressure has not been considered. If variations in
T and P along the capillary are significant, the situation is considerably complicated, as is explained in Section 8.6. For example the
Bagley plots are no longer straight lines.

FLOW IN CAPILLARIES. SLITS AND DIES

323

8.5.3 Rheological Significance of the Entrance Pressure Drop

For Newtonian fluids the entrance pressure drop is small, and the
larger values observed for polymeric liquids are therefore attributed
to the viscoelasticity of these liquids. The time-temperature superposition concept described in Section 2.12 suggests that a plot of
~Pends versus O"w should be independent of temperature, and this
has been observed [46].
Philippoff and Gaskins [47] proposed that the excess pressure
drop due to elasticity is related to the viscometric functions and to
the recoverable shear. However, when the ratio of the reservoir
radius to that of the capillary is large, as is the usual case, there will
be a high degree of stretching along streamlines as the melt flows
into the capillary. This means that the kinematics is very different
from that for viscometric flow, and except for weakly elastic materials, this relationship is not valid.
Noting the high degree of tensile extension (stretching along
streamlines) that occurs at the entrance to a capillary. Cogswell
[48,49] has derived an equation for calculating an apparent extensional viscosity using entrance pressure drop data. He assumes that
this pressure drop has two components, one due to shear and one
to extension. He further assumes that the extensional strain is
sufficiently large that the stored elastic energy resulting from
stretching is approximately equal everywhere to some maximum
value. Thus, the velocity distribution is governed by the viscous
resistance to flow and is calculated using the power law viscosity
model. The flow pattern is unknown, but Cogswell assumes the
streamlines will be those corresponding to the minimum pressure
drop. He thus derives the following formulas:
(8-54)

(8-55)

TfEC =

9{n + 1)2{~Pent)2
32 TfAyl

(8-56)

324

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

where iA is an average Hencky strain rate, and 'Y'JEC is an average


extensional viscosity.
Laun and Schuch [46] used Cogswell's equations together with
their data for several polymers to calculate values of 'Y'J EO and they
compared these with values of the true extensional viscosity determined using elaborate extensional rheometers. While there was
agreement for some resins over narrow ranges of strain rate, for
other resins and strain rate ranges there was no quantitative correlation between the two functions. However, in comparisons of two
resins, the values of IlPent (for a fixed value of Q and YA) were
found to give the same ranking of the materials as the true values of
'Y'JE at the value of (FE calculated from Equation 8-55.
Binding [50] has made a more elaborate analysis of the flow at a
flat entrance to a capillary. Assuming the fluid will adopt a flow
pattern that corresponds to the minimum energy dissipation, he
concludes that in extension thickening materials ('Y'J E increases with
i) vortices play the role of relieving the stresses that would result
from a sharp contraction of the flow. In this case the dominant
mechanism of dissipation is shear, and the pressure drop is governed by the viscosity. For extension-thinning materials ('Y'JE decreases with i), however, the dominant mechanism of dissipation is
extensional flow, and the entrance pressure drop is more closely
related to the extensional viscosity. There has been no experimental
verification of these phenomena, although White et al. [37,38] have
inferred from their experimental observations that there is a relationship between the occurrence of entrance vortices in slit flow
and the relative levels of shear and extensional stresses.
Ramanathan et al. [51] found the entrance pressure drop to be a
useful indication of the influence of plasticizer on the extrudate
quality of PVc.
8.6 CAPILLARY RHEOMETERS

The simplest and most ubiquitous type of melt rheometer is the


capillary rheometer. 1 While it is not very versatile in its capabilities,
i ASTM Standard Test Method D3835 involves the use of a capillary rheometer to characterize molten thermoplastics.

FLOW IN CAPILLARIES, SLITS AND DIES

325

its popularity makes it important for the practitioner to understand


its proper use and limitations.
In its simplest configuration, the capillary rheometer consists of a
small tube through which melt is made to flow, either by means of
an imposed pressure or a piston moving at a fixed speed. The
quantities normally measured are the flow rate, Q, and the driving
pressure, Pd' If the flow is generated by a moving piston, it is
usually the piston force, F d , that is measured, and this is related to
Pd as follows:
(8-57)
where Rb is the radius of the barrel or reservoir. Alternatively, Pd
can be measured by mounting a pressure transducer directly in the
barrel.
Capillary rheometers are used primarily to determine the viscosity in the shear rate range of 5 to 1000 s - \ although very long
capillaries have occasionally been used to extend the range to much
lower shear rates. To calculate the viscosity, it is necessary to know
the wall shear stress and the wall shear rate, and it is therefore
necessary to have reliable techniques for evaluating these basic
rheological quantities on the basis of experimental data. The wall
shear stress is related to the driving pressure by Equation 8-48:
Pd

O"w

-2(-L-/-R-+-e-)

(8-48)

We recall that "e" is the Bagley end correction, which is determined by one of the procedures described in Section 8.5.2. Also
discussed in that section is the possible rheological significance of
the entrance pressure drop.
It then remains to determine the wall shear rate. The general
method involves the determination of the Rabinowitch correction,
b, by use of Equation 8-20:
(8-20a)

326

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

where:
(8-20b)

We recall that l'A is the apparent wall shear rate defined by


Equation 8-13:
(8-13)

The determination of the Rabinowitch correction is a tedious and


time consuming procedure, and an alternative that usually gives a
good approximation of the true viscosity curve is the Schiimmer
technique described in Section 8.2.5. According to this procedure, a
curve of log( 7'1A) versus log( YA) can be shifted horizontally to yield
the true viscosity curve. In terms of equations, the shifting procedure corresponds to:
(8-21)

where x* is the shift factor and y* is the shifted shear rate. Then
the true viscosity at this shear rate is approximately (from Equation
8-22):

Schiimmer et al. [3,4] propose that with x* set equal to 0.83, the
viscosity determined in this way will differ from the true value at y*
by no more than 3%.
The principal sources of error in the use of a capillary rheometer
are:
1. Nonuniform temperature due to viscous heating or inadequate
heat transfer
2. Effect of pressure on viscosity
3. Wall slip or unsteady flow due to oscillating entry streamlines

Viscous heating always occurs when a viscous fluid is deformed, and


this will result in a temperature gradient in the fluid. This problem
is discussed by Dealy [52] who gives a method for estimating the

FLOW IN CAPILLARIES. SLITS AND DIES

327

magnitude of the difference between the wall temperature and the


maximum temperature in the melt. This difference becomes larger
as the viscosity, shear rate or radius is increased.
A pressure gradient is an essential element of capillary flow, and
since the viscosity is a function of pressure, the viscosity will vary
along the capillary. The resulting error becomes greater with a long
capillary, a high flow rate, a high viscosity and a small capillary
radius. This will lead to curvature of the Bagley plot of Pd versu~
LID. The usual procedure is to ignore the pressure effect and
associate the viscosity with the ambient pressure. However, an
alternative and more precise procedure is to take the pressure
dependency of the viscosity into account explicitly when analyzing
the data [46,53].
Wall slip is discussed in Section 8.2.6, and when this occurs it
renders the standard capillary flow equations invalid. If the velocity
profile remains axially symmetric, however, the equations presented
in that section can be used to determine the slip velocity and the
true value of Yw so that the viscosity can still be determined. The
difficulty is in knowing when slip is occurring. One effect of slip is a
small change in the slope of the flow curve (Iog(o-w) versus loge YA'
However, this may be difficult to detect with any certainty. Another
effect of slip is a change in the appearance of the extrudate, which
is usually observed to lose its surface gloss when slip occurs.
Probably the best guide is that slip is often found to occur when the
wall shear stress is a little above 0.1 MPa. If the extrudate exhibits
gross distortion, it probably means that the flow pattern in the
rheometer is quite complex so that the equations for calculating the
slip velocity and Yw are no longer valid.
Many types of capillary rheometer have been described in the
scientific literature, and commercial versions are available in a wide
range of degrees of sophistication and prices [52,54]. Appendix E
lists many manufacturers.
The simplest and most widely used of all melt testers is the
"extrusion plastometer" or "melt indexer." This is a primitive
capillary instrument in which the force is applied to the piston by
placing a weight on a platform attached to its upper end. ASTM 2

2American

Society for Testing and Materials, Philadelphia, PA.

328

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Standard Test Method D1238 is based on a specific embodiment of


this concept. This standard method prescribes the exact geometry
of the barrel and die and a choice of several standard temperatures
and driving weights for various materials. Condition "E" is the one
most used with polyethylene and involves a test temperature of
190C and a total load of 2.16 kg. The standard procedure involves
the measurement of the mass of polymer that is extruded in 10
minutes. The "flow rate," expressed as grams/l0 minutes (or
dg/min) is the number reported. When condition E is used, this
quantity is commonly called the "melt index" or MI. Obviously, the
MI value increases as the viscosity decreases, i.e., a "high MI" resin
has a low viscosity.
We note that while the quantities measured or specified can be
used to calculate the driving pressure and the apparent wall shear
rate, it is not possible to calculate the end correction, e, or the
Rabinowitch correction, b, so the true viscosity cannot be calculated. Furthermore, the L / D specified in the standard test method
is only about 4, which means that the entrance pressure drop will
constitute a significant part of the total pressure drop. A variation
in resin formulation can affect both the entrance pressure drop and
the viscosity, and the change in "flow rate" will reflect both of these
effects.
Other deficiencies of the melt indexer as a rheometer have been
enumerated [52], and the ASTM Standard Test Method warns the
user regarding the fundamental significance of the results. Even if
the test could be relied upon to yield a good approximation of the
viscosity at one shear stress, it still would have the limitation of
being a single-point characterization, which provides no information about the shape of the viscosity curve. Measuring the flow rates
at two or more loads, however, does provide some information
regarding the shape of the curve.
In any event, the "flow rate" or "melt index" is normally the only
specification of melt consistency that is provided by resin manufacturers. Furthermore, standard test methods very similar to ASTM
D1238 have been adopted by virtually every industrial nation, and it
is the rheological property nearly always cited for commercial
thermoplastic resins. For this reason, it has been selected by Saini
and Shenoy [55-58] as a normalizing parameter to produce master
curves of viscosity versus shear rate. In particular, they plot (7] X

FLOW IN CAPILLARIES, SLITS AND DIES

329

MI) versus (y JMI) for a generic polymer, say polystyrene, and


suggest that such a master curve is valid for any grade of that
polymer. They have published such master curves for polyolefins
[55, 56], polystyrene [55], engineering thermoplastics [57], and specialty polymers [58].
Testers designed in accordance with D1238 are offered by a
number of instrument manufacturers, most of whom are listed in
Appendix E. A detailed discussion of the test method and descriptions of many commercial melt indexers are given by Dealy [52].
Villemaire and Agassant [59] have described a capillary rheometer
in which a controlled amount of preshearing can be carried out in
concentric cylinder fixtures prior to flowing through the capillary. In
this way, the effects of shear history on viscosity can be studied. 3

8.7 FLOW IN CONVERGING CHANNELS


8.7.1 The Lubrication Approximation

For rectilinear flow in a straight channel, the fully-developed velocity profile and the pressure drop depend only on the viscosity
function in the range from y = 0 to y = Yw, and not on any other
rheological property. In classical (Newtonian) fluid mechanics, there
is an approximation technique that permits the assumption of the
fully developed velocity profile even when the cross section is not
constant in area. This is the "lubrication approximation," and is
valid when the Reynolds number is very small and the rate of
change of the size of the channel is also small. The requirement
that the Reynolds number be small ensures that the inertia (acceleration) terms in the equations of motion are negligible, and this
requirement is always satisfied for molten polymers because of their
high viscosity. The requirement concerning the change in cross
section dimension (dRjdz for a circular channel) be small ensures
that the viscous stress term in the Navier-Stokes equation will
approximate that for fully developed flow.

3A commercial version of this instrument, the "Rheoplast," is available from Metrilec, whose
address is given in Appendix E.

330

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

For a viscoelastic fluid, the latter requirement must be modified


to take into account the dependence of the viscous stress on
previous deformations of a fluid element. Another way of saying
this is that in order for the lubrication approximation to be valid,
the cross section must change sufficiently slowly that there is no
build up of tensile stress. A measure of the magnitude of the
extensional strain rate due to the convergence of the flow is the
axial gradient of the average axial velocity, dV/ dz, and a dimensionless group can be formed by multiplying this by a characteristic
relaxation time for the fluid, A. This group is a measure of the
extent to which fluid elasticity will play a significant role in the
motion and is clearly a Deborah number. Thus, the criterion for the
validity of the lubrication approximation is:
dV
A- 1
dz

(8-58)

For flow in a converging, circular channel, this is equivalent to:


(8-59)
Pearson [60] provides a more thorough discussion of the use of the
lubrication approximation in the simulation of melt processing
operations.
Using the lubrication approximation for flow through the conical
channel shown in Figure 8-9, the pressure drop for a power law
fluid is found to be:
2K
1 + 3n nyn [1 _(R
-dP _1 )3n]
- 3n tan a ( 4n ) A
Ro

(8-60)

where:
.

YA

4Q
- R3
7r

(8-61)

FLOW IN CAPILLARIES, SLITS AND DIES

331

Figure 8-9. Conical converging channel, showing the meaning of the symbols used in
Equation 8-60.

As a becomes smaller, the magnitude of !1P increases, because the


tapered section is longer. However, it is still often desirable to use a
small angle in designing a die, as it can reduce the severity of
extrudate distortion that can occur at high flow rates (see Section
8.9). When the convergence angle is not small, experimental data
indicate that the pressure drop is independent of convergence angle
at high flow rates, and this reflects the fact that the lubrication
approximation is no longer valid when a is not small.
The lubrication approximation has also been used to estimate the
velocity distribution for large entry angles. For example, Cogswell's
method for determining an apparent extensional viscosity using the
capillary entrance pressure (see Section 8.5.3) makes use of this
technique, and Gibson [61] has proposed the use of Cogswell's
model for converging channel flow. Another approach is to assume
that the velocity profile is governed by the viscosity function but
that the first normal stress difference affects the pressure drop
[62,63]. This implies that the Weissenberg number (QA/Rf) is not
negligible, while the Deborah number continues to be small enough
so that tensile stresses can be neglected. This could occur, for
example, when Q is large but a is small, and the results of these
analyses predict the independence of !1P on a at large Q that is
observed experimentally [62,63].

332

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

8.7.2 Industrial Die Design

When the Deborah number (Equation 8-58) is not small, the flow
can only be reliably modelled by a numerical technique that incorporates a nonlinear viscoelastic constitutive equation. However,
because of the complexities of this procedure, most die design
formulas used in the plastics industry are based on the use of the
lubrication approximation together with the power law model
[64-66].
8.8 EXTRUDATE SWELL

At the exit of a die, the emerging stream of melt rapidly increases


in cross-sectional area, and this phenomenon is called die swell or
extrudate swell. If the flow channel is noncircular, the extrudate will
also undergo a change of shape. Such changes in the shape and size
of an extrudate are of practical importance in such processes as
profile extrusion, blow molding and injection molding. In addition,
swell has been used as a qualitative measure of melt elasticity for
quality control purposes.
The simplest type of swell is that which occurs at the exit of a
circular die, such as an orifice plate or a capillary. There is no
change of shape, and the swell can be described in terms of a single
"swell ratio," i.e., the ratio of the diameter of the extrudate to that
of the die lip. Laun and Schuch [46] have studied the effect of die
geometry on this ratio and found that a diverging die decreases the
swell, while a converging die increases it.
The variables commonly used to describe swell are the swell
ratio, for which we will use the symbol B, the fully developed wall
shear rate, Yw, the length to diameter ratio, LID, the temperature,
T, and the time, t, that has elapsed since the melt left the die.
Thus, for a given polymer, we can show the relevant variables as
follows:

B == DIDo

f(yw' LID, T, t)

(8-62)

This ratio is about 1.13 for a Newtonian fluid [67-70], but for
polymeric liquids it can be as high as 2 or even 3. This large swell
ratio is a manifestation of the molecular orientation that is gener-

FLOW IN CAPILLARIES, SLITS AND DIES

333

ated by the flow in the die. At the inlet from a larger reservoir,
streamlines converge rapidly, and this generates a high degree of
stretching along streamlines. In a short die or orifice, this leads to a
large swell at the exit, where the melt is free to deform in response
to the anisotropic stresses associated with the molecular orientation. These large extensional stresses can be determined by means
of birefringence measurements, and plots of tensile stress along the
flow axis of a slit as a function of axial distance have been reported
by Brizitsky et al. [70].
However, if the entrance leads to a capillary, relaxation will lead
to a loss of the orientation generated at the entrance, and as the
capillary is lengthened the degree of swell is reduced [71]. At the
same time, however, shearing in the die produces a moderate
degree of axial orientation, and this will lead to a certain degree of
extrudate swell. For a long capillary, therefore, the swell becomes
independent of length and is a reflection only of the shear-induced
orientation generated in the capillary. The effect of LjD on swell
ratio is shown clearly in the graph published by Han et al. [71] and
shown in Figure 8-10. This graph also shows that the swell ratio
increases with the wall shear rate.
2.6,..------:-------------------,
co 2.5

~ 2.4
a:
-l

ul

2.3

2.2

f-

a:

2.1

f-

i';'j 2.0
1.9L-L-~~~~~~~~~~~~~~~~~~

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21
L/D

Figure 8-10. Extrudate swell ratio versus LjD for HDPE at 180C for three shear rates.
From top to bottom, these are, in s -1: 700, 400 and 200. Adapted from Ref. 71. Copyright
1970 by The Society of Rheology. Reprinted by permission of John Wiley & Sons, Inc.

334

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Various experimental methods for measuring extrudate swell


have been reviewed and compared by Graessley et al. [72], White
and Roman [73] and Utracki et al. [74]. For rapid evaluation for
quality control purposes, the diameter of the extrudate at some
fixed distance from the die can be measured directly by means of an
optical method. Die swell detectors of this type are made by
Monsanto and Brabender. Kurtz et al. [75] have described an
automatic system for continuous on-line use.
The most commonly used laboratory methods involve either
extruding into air and measuring the diameter of the frozen extrudate or measuring the diameter of such an extrudate after annealing to permit the diameter to reach its fully swollen value. However,
the most reliable method is to extrude into a hot oil having the
same density as the melt. In this way, sagging under the influence of
gravity can be prevented while the extrudate is allowed to reach its
equilibrium swell value.
Because swell is a viscoelastic phenomenon, it depends on time
and temperature. Therefore, the time dependency can only be
studied experimentally by extruding into an oil having the same
density and temperature as the melt in the capillary. This technique
has been used to measure swell as a function of time for a capillary
[74], a straight annular die [76] and for several converging and
diverging dies [77]. For HDPE at 170C about 75% of the swell
occurred in the first few seconds, while the ultimate swell, Boo, was
only reached after several minutes. For a polypropylene at 190C,
only 50% of the swell occurred in the first few seconds while more
than 10 minutes was required to reach the equilibrium swell.
For a given resin, if attention is restricted to the ultimate swell,
and only capillaries with large L / D are used, the independent
variables of Equation 8-62 are reduced to 2:
(8-63)
Han [42, p. 117] has observed that if the ultimate swell is represented as a function of the wall shear stress rather than the shear
rate, the dependence on temperature is eliminated:

(8-64)

FLOW IN CAPILLARIES, SLITS AND DIES

335

Theoretical analyses of extrudate swell are of interest in two


contexts. First, it was at one time hoped tliat it might be possible to
determine Ni y) by making measurements of B( Yw)' as Nl is
difficult to measure by more direct techniques, especially at high
shear rates. However, swell is not at all a viscometric flow, and it is
now recognized that there is no simple relationship between swell
ratio and Nl [69]. In fact, computational studies suggest that
extrudate swelling is very sensitive to the shape of the extensional
stress growth curve [78].
The second application of a model of extrudate swell is in the
prediction of swell for a given die design and polymer. This would
make it possible, for example, to design a blow molding die to
produce a given degree of swell. This involves the use of a numerical computation scheme together with a nonlinear constitutive
equation to describe the viscoelastic behavior of the melt. While
much interesting work has been done [79,80], it has not yet produced reliable predictive models for industrial design purposes [78].
Because commercial extrusion dies usually have their own temperature control system, it is of practical interest to examine the
effect on the swell of a difference between the stock temperature
and the die wall temperature [81,82]. The results of such studies
cannot be generalized [82], but cooling the die wall is found to
substantially increase the swell of some thermoplastics [81]. These
effects cannot be observed using a capillary rheometer [81].
For noncircular dies [83,84] both the size and the shape of the
extrudate vary with time after extrusion. Stevenson et al. [85] have
described a scheme for classifying dimensional changes for complex
profile extrusions on the basis of size and shape factors. They also
describe methods for controlling size and shape by manipulation of
line speed, stock temperature and die temperature.
The dependence of swell on the molecular weight distribution is
of great importance in the plastics industry, as it is often desired to
produce a resin having prescribed swelling characteristics. While it
has been generally accepted that increasing the molecular weight
increases swell at constant flow rate, no simple generalizations
about the effect of molecular structure on swell are possible because of the many parameters necessary to describe adequately this
structure. Even for a perfectly linear polymer, many moments of
the molecular weight distribution may be required to provide an

336

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

adequate description of the distribution for a commercial polyolefin. In the case of polyethylene, some degree of long chain
branching is often present, even in a "linear" polymer, and this
complicates the picture further. For these reasons, attempts to
correlate swell with MWD using only Mw and Mn have led to
conflicting conclusions [86, 87].
Koopmans [88] has helped to resolve this conflict by studying the
short-time capillary swell of a series of blends of two HDPE's, one
having a much higher value of M w than the other. He found that
the dominant variable for swell was the molecular weight of the
high molecular weight component.

8.9 EXTRUDATE DISTORTION

In the previous section we saw that the extrudate from a capillary


or die can undergo a large change in the size and shape of its cross
section. Another phenomenon that can pose an even greater problem in melt processing is extrudate distortion, by which is meant a
variation in shape along the length of the extrudate. This effect can
range in intensity from a loss of gloss to gross distortion and is the
factor that limits the production rate in certain processes such as
the blown film processing of LLDPE. A term sometimes used to
describe the various extrudate distortion effects is "melt fracture."
Many phenomenological observations of extrudate distortion have
been made, for example by Ballenger et al. [89] and by Oyanagi
[90]. Extensive reviews of published observations and measurements
have been published by White [91] and by Petrie and Denn [92].
While the specific nature of the phenomena observed varies considerably from one polymer to another, there are a few nearly universal features. One is that the first onset of noticeable distortion
occurs at a critical wall shear stress in the neighborhood of 0.1
MPa. Another nearly universal observation is that over certain
ranges of flow rate, the extrudate distortion takes the form of a
more or less regular, helical distortion sometimes having the appearance of a screw thread. There is also a type of extrudate
distortion called "palm tree," which occurs with filled elastometers
and aluminum. A fairly smooth extrudate is periodically interrupted
by short rough section of much larger diameter.

FLOW IN CAPILLARIES, SLITS AND DIES

337

The factors that influence the type and degree of extrudate


distortion include:
Chemical nature of polymer
Molecular weight and its distribution
Branching
Entrance geometry
Length to diameter ratio of the capillary
Material of construction of the capillary
Temperature
Flow rate
Narrowing the molecular weight distribution or increasing the average molecular weight usually increase the severity of the effect. If
the criterion for the onset of extrudate distortion is expressed in
terms of a critical shear stress, this quantity is found to be independent of temperature and flow rate.
There is as yet no general understanding of the causes of extrudate distortion, and we must therefore begin this discussion by
simply describing the phenomena that have been observed in various materials. We will restrict our attention for the moment to
capillary flow, which is the situation most studied to date. Most
observations have been made using an apparatus in which the flow
is generated by a piston moving at constant speed in a reservoir
above the capillary. Thus, the rate of volume displacement in the
reservoir is constant with time. As we shall see, this does not
necessarily imply a constant flow rate in the capillary because of the
possible combined effects of pressure fluctuations and melt compressibility.
8.9.1 Surface Melt Fracture: Sharkskin

At low wall shear stresses, the extrudate is always smooth and


glossy when the resin is a single phase material and well above its
melting point. As the flow rate is increased, however, a point is
reached where molten thermoplastics begin to exhibit a surface
roughness, which is sometimes called "matte," "loss of gloss" or
"surface melt fracture." In its more severe form, this phenomenon

338

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

is often called "sharkskin." Close observation of the extrudate


usually reveals that the distortions have the appearance of a fine
helical screw thread.
By using a technique involving carbon black filled tracer resin
and the slicing of extrudate, Bergem [93] showed that these screw
threads originate at the exit of the die. Cogswell [94] has suggested
that this type of distortion results from the abrupt change in the
boundary condition at the exit of the die, which causes a high
degree of stretching in the surface layer of the melt as it leaves the
die. This stretching produces tensile stresses that can exceed the
strength of the melt, leading to surface fracture. Later careful
studies of extrudates of linear polyethylenes, however, suggest that
the onset of this surface fracture is accompanied by the occurrence
of wall slip in the capillary [5,6]. Thus, while this type of distortion
is clearly an exit effect, it may be associated with slip flow in the
capillary. As shown in Section 8.2.6, when slip starts to occur,
its presence is reflected in a change in the slope of a plot of O"w
versus YA'
8.9.2 Oscillatory Flow in Linear Polymers

For several linear polymers, particularly polyethylene, as the flow


rate is further increased, a regime is reached in which the flow
rate and pressure oscillate between upper and lower bounds.
Vinogradov [95,96] has called this the "spurt effect." In this regime,
the extrudate usually consists of regularly alternating zones of
relatively rough and relatively smooth extrudate. It has been hypothesized that these correspond to periods during which the melt
alternately sticks and slips at the wall of the capillary.
This implies a variation in the flow rate with time. If the piston
moves at constant speed, and the melt were truly incompressible,
an oscillation in the flow rate would be impossible, but Lupton and
Regester [97] have shown how even the small degree of compressibility that is present in melts, together with the presence of large
pressure oscillations, can account for the flow rate variations.
In the case of some linear polyethylene resins, it has been
observed [89,97] that in the "rough" zones the extrudate has a
generally helical or rope-like appearance while on a smaller but still

FLOW IN CAPILLARIES. SLITS AND DIES

339

visible scale the surface is smooth. Bergem [93] and Piau et al. [36]
have shown that this helical pattern is directly related to a swirling
motion of the melt at the entrance to the capillary. The association
of high flow rate extrudate distortion in linear polyethylene with
entrance effects is supported by the observation that the die entry
geometry has a significant effect on the intensity of the effect, while
die length has little effect.
Ramamurthy [5] found that for both HDPE and LLDPE, surface
melt fracture and a small reduction in the slope of the flow curve
occur at a wall shear stress of about 0.145 MPa, while the transition
to oscillatory flow occurs at o"w = 0.43 MPa. Kalika and Denn [6]
noted that in the oscillating flow phase of the phenomenon, the
extrudate exhibits alternately gloss and severe sharkskin. Furthermore, they reported that the glossy segments correspond to the
part of the cycle during which the pressure was falling, while
Ramamurthy [5] said he observed the opposite situation.
Kalika and Denn [6] noted that the period of the oscillation
decreased as the piston descended, i.e., as the reservoir volume
decreased. This is in accord with the analysis of compressibility
effects published by Lupton and Regester [97].

8.9.3 Gross Melt Fracture

Further increases in flow rate above the range over which oscillating pressure is observed lead to an intense random distortion,
which Kalika and Denn [6] describe as "waviness." In this regime,
calculation of a slip velocity using Equation 8-29 gave results that
were double those at the upper end of the oscillatory flow regime.
This suggests either a second mode of slip or a flow pattern in
the capillary that is more complex than that assumed in deriving
Equation 8-29.
For polystyrene, an initially smooth extrudate is observed to show
a helical screw thread distortion at o"w ::::: 0.1 MPa, with the value
decreasing as Mw increases [89,98]. At higher flow rates, gross
distortion is observed, with no discontinuity in the flow curve. The
behavior for polypropylene has been reported to be similar to that
of polystyrene [89,99].

340

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

8.9.4 Role of Slip in Melt Fracture

Ramamurthy [5] carried out an elaborate research program to study


the role of slip in the generation of extrudate distortion. He used
the procedure described in Section 8.2.6 (Plot of YA versus l/R for
a fixed O'w) to show that slip does occur in the case of linear
polyethylenes and to calculate the slip velocity. Kalika and Denn [6]
have also reported slip velocities for LLDPE.
Ramamurthy also studied the effect of changing the material of
construction of the capillary, and he found the effect to be a weak
one. However, in studies using an extruder-fed film die he found a
marked effect of material of construction. This effect was strongly
time dependent, with a significant induction time required to achieve
the full effect. This time dependence was very small for chromeplated steel, but for copper the severity of the distortion increased
with time, and it decreased for stainless steel and aluminum [100].
For certain brasses, slip and extrudate distortion were eliminated at
all observable flow rates after a long induction time [101]. This led
Ramamurthy to hypothesize that slip is primarily a physico-chemical effect and not simply a mechanical friction effect. This hypothesis is supported by the observation that hydrostatic pressure has
little effect on extrudate distortion [6,7].
Ramamurthy [5] also studied the mechanism by which fluorocarbon additives suppress extrudate distortion and proposed that these
materials, usually considered to be lubricants, act as adhesion
promoters! Again, a lengthy induction time is necessary before the
effect is observed, which means that it normally would not show up
in a capillary rheometer. Ramamurthy has proposed [102] that
sharkskin can be eliminated without a reduction in flow rate by a
judicious selection of die material combined with the use of such an
additive.
Piau et al. [36] made careful observations of the extrudate as well
as the entrance flow of branched and linear silicone fluids. They
used both orifice dies and short capillaries and observed all the
classical extrudate distortions: mild roughness, sharkskin, helical
patterns, and gross melt fracture. They concluded that while mild
roughness and sharkskin are exit effects, the more severe types of
distortion are always direct results of entrance flow disturbances.
Finally, comparing the results for dies made of steel and of PTFE,

FLOW IN CAPILLARIES, SLITS AND DIES

341

they concluded that wall slip has little, if anything, to do with


extrudate distortion. They argue that Ramamurthy's calculations
of wall slip values could have been artifacts associated with the
entrance pressure drop.

8.9.5 Gross Melt Fracture Without Oscillations

For LDPE the first appearance of surface roughness is followed


almost immediately by the occurrence of gross melt fracture, at a
wall shear stress of about 0.13 MPa. For LDPE and for polystyrene,
when corner entrance vortices are present, gross melt fracture is
associated with a complex entrance flow in which the central
"funnel" that is formed in the center of the ring vortex breaks at
random intervals, with melt suddenly spilling into the capillary from
random angular positions in the vortex. In LDPE, the extrudate
distortion decreases with the length of the capillary, presumably
due to the time allowed to "heal" the entrance fracture. For this
polymer, the flow curve does not exhibit a discontinuity, although
Ramamurthy [5] has reported a small reduction in the slope at a
wall shear stress of about 0.13 MPa.
Yamane and White [103] observed the extrudate appearance for
a large number of polyethylenes of all types. While their findings
were in general accord with the above discussion, they found
significant variations in detailed behavior among resins of the same
nominal type. This appears to result, at least in part, from
the frequent presence of some degree of branching in "linear"
polyethylenes.
REFERENCES
1. K. Walters, Rheometry, Chapman and Hall, London, 1975.
2. R. W. Whorlow, Rheological Techniques, Ellis Norwood, distributed by John
Wiley & Sons, New York, 1979.
3. P. Schiimmer and R. H. Worthoff, Chern. Eng. Sci. 33:759 (1978).
4. H. Chmiel and P. Schiimmer, Chemie-Ing.-Techn. 43:1257 (1971).
5. A. V. Ramamurthy, J. Rheo!. 30:337 (1986).
6. D. S. Kalika and M. M. Denn, J. Rheol. 31:815 (1987).
7. H. M. Laun, Rheol. Acta 22:171 (1983).

342

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

8. A. S. Lodge, "Normal stress differences from hole pressure measurements,"


Chapter 11 of Rheological Measurement, edited by A. A. Collyer and D. W.
Clegg, Elsevier Applied Science, London & NY, 1988.
9. R. I. Tanner and A. C. Pipkin, Trans. Soc. Rheol. 13:471 (1979).
10. K. Higashitani and W. G. Pritchard, Trans. Soc. Rheol. 16:687 (1972).
11. R. D. Pike and D. G. Baird, 1. Non-Newt. Fl. Mech. 16:211 (1984).
12. D. G. Baird, M. D. Read and R. D. Pike, Polym. Eng. Sci. 26:225 (1986).
13. D. G. Baird, Trans. Soc. Rheol. 19:147 (1975).
14. D. G. Baird, 1. Polym. Sci. 20:3155 (1976).
15. F. Sugeng, N. Ph an-Thien and R. I. Tanner, 1. Rheol. 32:215 (1988).
16. D. G. Baird, M. D. Read and J. N. Reddy, 1. Rheol. 32:621 (1988).
17. A. S. Lodge and L. de Vargas, Rheol. Acta 22:151 (1983).
18. R. Srinivasan and B. A. Finlayson, 1. Non-Newt. Fl. Mech. 27:1 (1988).
19. C. D. Han, Trans. Soc. Rheol. 22:171 (1983).
20. J. M. Davis, J. F. Hutton and K. Walters, 1. Phys. (D) 6:2259 (1973).
21. C. D. Han, Trans. Soc. Rheol. 18:163 (1974).
22. M. Gottlieb and R. B. Bird, Ind. Eng. Chem. Fund. 18:357 (1979).
23. C. D. Han and L. H. Drexler, Trans. Soc. Rheol. 17:659 (1973).
24. K. R. Reddy and R. I. Tanner, 1. Rheol. 22:66 (1978).
25. N. Y. Tuna and B. A. Finlayson, 1. Rheol. 28:79 (1984).
26. J. Vlachopoulos and E. Mitsoulis, 1. Polym. Eng. 5:173 (1985).
27. D. V. Boger and M. M. Denn, 1. Non-Newt. Fl. Mech. 6:163 (1980).
28. C. D. Han, "Slit Rheometry," Chapter 2 of Rheological Measurement, edited
by A. A. Collyer and D. W. Clegg, Elsevier Applied Science, London & NY,
1988.
29. C. Rauwendaal and F. Fernandez, Polym. Eng. Sci. 25:765 (1985).
30. N. Y. Tuna and B. A. Finlayson, J. Rheol. 32:285 (1988).
31. W. Kozicki, C. H. Chou and C. Tiu, Chem. Eng. Sci. 21:665 (1966).
32. C. Miller, Ind. Eng. Chem. Fundam. 11:524 (1972).
33. T.-J. Liu and C.-N. Hong, Polym. Eng. Sci. 28:1559 (1988).
34. J. L. White and A. Kondo, 1. Non-Newt. Fl. Mech. 3:41 (1977).
35. J. M. Piau, N. El Kissi and B. Tremblay, 1. Non-Newt. Fl. Mech. 30:197
(1988).
36. J. M. Piau, N. El Kissi and B. Tremblay, "Influence of upstream instabilities
and wall slip on melt fracture and sharkskin phenomena during silicone
extrusion through orifice dies," to be published, 1990.
37. S. A. White and D. G. Baird, 1. Non-Newt. Fl. Mech. 29:245 (1988).
38. S. A. White, A. D. Gotsis and D. G. Baird, 1. Non-Newt. Fl. Mech. 24:121
(1987).
39. C. D. Han and L. H. Drexler, 1. Appl. Polym. Sci. 17:2369 (1973).
40. S. T. E. Aldhouse, M. R. Mackley and I. P. T. Moore, 1. Non-Newt. Fl.
Mech. 21:359 (1986).
41. M. R. Mackley and I. P. T. Moore, 1. Non-Newt. Fl. Mech. 21:337 (1986).
42. C. D. Han, Rheology in Polymer Processing, Academic Press, New York, 1976.
43. E. B. Bagley, J. Appl. Phys. 28:624 (1957).

FLOW IN CAPILLARIES. SLITS AND DIES

44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
65.
66.
67.
68.
69.
70.
71.
72.
73.
74.
75.
76.
77.

343

H. L. Weissberg, Phys. Fl. 5:1033 (1962).


F. N. Cogswell, Polymer Melt Rheology, John Wiley & Sons, New York, 1981.
H. M. Laun and H. Schuch, J. Rheol. 33:119 (1989).
W. Philippoff and F. H. Gaskins, Trans. Soc. Rheol. 2:263 (1958).
F. N. Cogswell, Trans. Soc. Rheo116:383 (1972).
F. N. Cogswell, Polym. Eng. Sci. 12:64 (1972).
D. M. Binding, J. Non-Newt. Fl. Mech. 27:173 (1988).
R. Ramanathan, D. G. Baird and L. G. Krauskoff, SPE Tech. Papers 3:773
(1989).
J. M. Dealy, Rheometers for Molten Plastics, Van Nostrand Reinhold, New
York,1982.
R. C. Penwell, R. S. Porter and S. Middleman, J. Polym Sci. A-2 9:731
(1971).
G. H. France, Chapter 7 of Rheological Measurement, edited by A. A. Collyer
and D. W. Clegg, Elsevier Applied Science, London and New York, 1988.
A. V. Shenoy, S. Chattopadhyay, and V. M. Nadkami, Rheol. Acta 22:90
(1983).
D. R. Saini and A. V. Shenoy, Eur. Polym J. 19:811 (1983).
A. V. Shenoy, D. R. Saini and V. M. Nadkami, Rheol. Acta 22:209 (1983).
D. R. Saini and A. V. Shenoy, J. Elastomers Plastics 17:189 (1985).
J. P. Villemaire and J. F. Agassant, Polym. Proc. Eng. 1:223 (1983-4).
J. R. A. Pearson, Mechanics of Polymer Processing, Elsevier Applied Science
Publishers, New York, 1985.
A. G. Gibson, Chapter 3 of Rheological Measurement, edited by A. A. Collyer
and D. W. Clegg, Elsevier Applied Science, London and New York, 1988.
R. L. Boles, H. L. Davis and D. C. Bogue, Polym. Eng. Sci. 10:24 (1970).
T. H. Kwon, S. F. Shen and K K Wang, Polym. Eng. Sci. 26:214 (1986).
W. Michaeli, Extrusion Dies, Hanser, Munich and New York, 1984.
Z. Tadmor and C. G. Gogos, Principles of Polymer Processing, John Wiley &
Sons, New York, 1979 (Chapter 13).
T.-J. Liu, C.-N. Hong and K-C. Chen, Polym. Eng. Sci. 28:1517 (1988).
J. Gavins and M. Modan, Phys. Fl. 10:487 (1967).
R. I. Tanner, Appl. Polym. Symp. 20:201 (1973).
K Reddy and R. I. Tanner, J. Rheol. 22:661 (1978).
V. I. Brizitsky, G. V. Vinogradov, A. I. Isayev and Y. Y. Podolsky, J. Appl.
Polym. Sci. 27:751 (1978).
C. D. Han, M. Charles and W. Philippoff, Trans. Soc. Rheol. 14:393 (1970).
W. W. Graessley, S. D. Glasscock and R. L. Crawley, Trans. Soc. Rheol. 14:19
(1970).
J. L. White and J. F. Roman, J. Appl. Polym. Sci. 20:1005 (1976).
L. A. Utracki, Z. Bakerdjian and M. R. Kamal, J. Appl. Polym. Sci. 19:481
(1975).
S. J. Kurtz, T. A. De Rossett and M. T. Shaw, U.S. Patent 4,449,395, (1984).
A. Garcia-Rejon and J. M. Dealy, Polym. Eng. Sci. 22:158 (1982).
N. Orbey and J. M. Dealy, Polym. Eng. Sci. 24:511 (1984).

344

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

78. R. I. Tanner, "Recoverable Elastic Strain and Swelling Ratio," Chapter 4 of


Rheological Measurement, edited by A. A. Collyer and D. W. Clegg, Elsevier
Applied Science, London and New York, 1988.
79. M. J. Crochet, A. R. Davies and K. Walters, Numerical Simulation of
Non-Newtonian Flow, Elsevier, Amsterdam 1984.
80. R. I. Tanner, J. Rheol. 32:673 (1988).
81. A. M. Henderson and A. Rudin, J. Appl. Polym. Sci. 31:353 (1986).
82. B. Yang and L. J. Lee, Polym. Eng. Sci. 27:1079 (1987).
83. J. L. White and D. Huang, Polym. Eng. Sci. 21:1101 (1981).
84. D. C. Huang and J. L. White, Polym. Eng. Sci. 19:609 (1979).
85. J. F. Stevenson, L. J. Lee and R. M. Griffith, Polym. Eng. Sci. 26:233 (1986).
86. M. G. Rogers, J. Appl. Polym. Sci. 14:1679 (1970).
87. R. A. Mendelson and F. L. Finger, J. Appl. Polym. Sci. 19:1061 (1975).
88. R. J. Koopmans, J. Polym. Sci. A 26:1157 (1988).
89. T. F. Ballenger, I.-Jen Chen, J. W. Crowder, G. E. Hagler, D. C. Bogue and
J. L. White, Trans. Soc. Rheol. 15:195 (1971).
90. Y. Oyanagi, Appl. Polym. Symp. 20:123 (1973).
91. J. L. White, Appl. Polym. Symp. 20:155 (1973).
92. C. J. S. Petrie and M. M. Denn, A. I. Ch.E. J. 22:209 (1976).
93. N. Bergem, Proc. VJIIth Int. Congr. Rheol., Gothenberg, 1976, p. 50.
94. N. F. Cogswell, Polymer Melt Rheology, John Wiley & Sons, New York, 1981,
p. 101.
95. G. V. Vinogradov, N. I. Insarova, B. B. Boiko and E. K. Borisenkova, Polym.
Eng. Sci. 12:323 (1972).
96. G. V. Vinogradov, Polymer 18:1275 (1977).
97. J. M. Lupton and R. W. Regester, Polym. Eng. Sci. 5:235 (1965).
98. R. S. Spencer and R. E. Dillon, J. Colloid Sci. 4:241 (1949).
99. M. Fujiyama and H. Awaya, J. Appl. Polym. Sci. 16:275 (1972).
100. A. V. Ramamurthy, U.S. Patent 4,552,712 (1985).
101. A. V. Ramamurthy, U.S. Patent 4,554, 120 (1985).
102. A. V. Ramamurthy, U.S. Patent 4,522,776 (1985).
103. H. Yamane and J. L. White, J. Rheol. Japan 15:131 (1987).

Chapter 9
Rhea-Optics and Molecular
Orientation

9.1 BASIC CONCEPTS: INTERACTION OF LIGHT AND MATTER

A beam of light may be thought of as an oscillating electric field


propagating through space. Matter is composed of electrically
charged electrons and nuclei, which are affected by electric fields.
Therefore, when a light beam encounters a material, its electric
field interacts with the charged components of the matter, and the
light beam is altered by this interaction. The type and magnitude of
the interaction with the light can be used to probe the state of the
matter. Application of optical techniques to deforming systems
("rheo-optics") allows the measurement of components of the
velocity vector and of the stress tensor.
Optical methods have a number of significant advantages over
mechanical methods of measurement [1]. They do not require
contact with, and consequent perturbation of, the deforming system. The light can be focused on a small volume element, and one
can therefore measure the state of this small volume element,
rather than the average state of the entire body. Also, the frequency response is high, permitting measurements on very small
time scales. By proper choice of wavelength it is sometimes possible
to isolate the behavior of one component of a mixture, or of one
structural element of a molecule.
Optical measurement of the velocity and its gradient is properly
the concern of fluid mechanics rather than of rheology. We give
here, therefore, only references in which the interested reader can
find detailed descriptions of techniques and apparatus. An extensive tabulation of papers and descriptions of techniques is given by
345

346

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Mackay and Boger [2]. Other references of interest are papers on


two-color laser Doppler velocimetry [3], laser speckle interferometry [4], and homodyne light scattering [5].
9.1.1 Refractive Index and Polarization

The simplest interaction of matter and light is the transmission of a


beam by an isotropic, uniform, nonabsorbing medium. The result of
the interaction is to decrease the velocity of the light in the medium
as compared to the velocity in a vacuum. The ratio of the velocity in
a vacuum compared to that in the medium is the familiar refractive
index n. In this instance n is a real scalar quantity.
In order to consider the interaction with more complex media it
is advantageous to have the incident beam in a known state of
"polarization." This can be done experimentally. Polarization refers
to the state of the electric field vector [6,7]. The vector is perpendicular to the direction of propagation z of the beam, and can be
resolved into components Ex and Ey along two mutually perpendicular directions x and y both of which are orthogonal to z. Each
component E; (i = x or y) oscillates at the frequency w of the light,
but has its own amplitude, A; and phase, 0;. (The phase angle 8
used in this chapter is not to be confused with the mechanical loss
angle 8 measured in dynamic mechanical experiments.) The time
and distance dependence of each component of the electric field
vector is then given by
E;

A; cos{w(t - nz/c)

+ oj

(9-1)

i=xory
The velocity v of the light in the medium is c/n, where c is the
velocity of light in a vacuum, and n is the index of refraction of the
medium.
Equation 9-1 is a solution of the classical equations of electromagnetic theory, derivations of which are to be found in the
literature [7]. The consequences of Equation 9-1 are appreciated
most readily by considering some special cases. At any given instant
of time, both Ex and Ey vary sinusoidally with linear distance z
through the medium. The wavelength A of the oscillation is 21TV / W.

RHEO-OPTICS AND MOLECULAR ORIENTATION

347

The instantaneous resultant electric field is the vector sum of the


two components Ex and E y Their ratio is the tangent of the angle
of the resultant vector with the x-coordinate axis. Now, if i>x and i>y
are equal, the two components of the field are said to be in-phase,
and their ratio is uniform and independent of time. In that case the
direction of the resultant electric vector remains fixed, and the light
is said to be "linearly polarized." This is also the case if the
difference of the two phase angles is just 180 (77' radians).
If, on the other hand, the phase angles are not equal, the
direction of the resultant vector changes with time. The common
term w(t - nz/c) can be eliminated from the two Equations 9-1 to
give a relationship between Ex and E y. This relation has the
mathematical form of the equation for an ellipse, and the light is
therefore said to be polarized elliptically. The eccentricity of the
ellipse, and the orientation of its axes with respect to the x and y
axes, depends on the difference i> of the phase angles, i> = i>y - i>x'
and on the ratio of Ax to A y' In the particular case i> = 77' /2 and
Ax = A y' the ellipse becomes a circle, and the light is then circularly polarized. These relationships are illustrated in Figure 9-1 for
the cases Ay = Ax and Ay = 2Ax'
Because it makes for a compact notation, and because it simplifies many mathematical manipulations, it has become conventional
in the optics literature to express Equation 9-1 using complex
variable notation, as follows:
0

Ex =Ax Re[exp{iw{t - nz/c) + ii>xl]

(9-2a)

Ay Re[ exp{iw{t - nz/c) + ii>y}]

(9-2b)

Ey

In these equations i is the square root of -1. Re means "the real


part of the complex expression in brackets," where
Re [exp(i cf>)]

Re [cos cf> + i sin cf>]

cos cf>

9.1.2 Absorption and Scattering

All materials absorb light of some frequency in the spectrum of


electromagnetic radiation. Absorption results in attenuation of the

348

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

~oo/
Il = -3"./4

Il = -"./2

Il = "./4

Il = "./2

-"./4

Il = 0

Il = 3"./4

Il=".

Il

(a)

\J 0 0 /
0 0 \J \

Il = -3"./4

Il = "./4

Il = -"./2

Il = "./2

Il = -"./4

1l=0

Il = "./4

Il=".

(b)

Figure 9-1. Polarization diagrams; plots of Ey versus Ex, as calculated from Equation 9-1,
for various values of the angle Il. (a) Ay = Ax, (b) Ay = 2Ax. Adapted from Ref. 7.
Copyright 1984 by John Wiley & Sons, Inc. Reprinted by permission.

intensity of the light beam as it passes through the medium.


Mathematically this attenuation can be represented by a complex
refractive index n in Equation 9-1, where
n

n' - in"

(9-3)

Substituting Equation 9-3 into 9-1, the expression for the amplitude E becomes
E

+ izn"/c) + io]

A Re[exp{iw(t - zn'/c

A exp{ -zwn"/c} Re[exp{iw(t - zn'/c)

+ io}]

RHEO-OPTICS AND MOLECULAR ORIENTATION

349

The last expression shows that the amplitude of the electric


vector and thus the intensity of the light beam decreases exponentially with the distance of passage through the absorbing medium.
The "extinction coefficient" for the rate of decrease of intensity is
proportional to nil, the imaginary component of the refractive
index.
Absorption of light is not the only possible cause for the existence of an imaginary component in the refractive index. Light is
reflected and refracted at the interface of two materials whose
refractive indices differ. Therefore, if a material is an inhomogeneous mixture, consisting of regions with different n's, the light
reflected and refracted from the multiple interfaces, and the interferences of these secondary beams, will cause scattering of light
into directions different from those of the incident beam. The
angular distribution of the intensity and polarization of the scattered light depend upon the size and structure of the scattering
centers. Scattering measurements to determine the structure of
scatterers have not been widely applied to molten polymers. We
refer the reader, therefore, to the literature for experimental results
and for references to the theories [8-10].

9.1.3 Anisotropic Media: Birefringence and Dichroism

In addition to making the refractive index a complex scalar, as we


have discussed, another extension of Equation 9-1 to more complex
media allows us to consider it as a tensor quantity. This extension
enables us to describe the transmission of light by anisotropic
media.
If the linear density of the electrically charged particles of matter
differs along various directions in a body, the interaction of the light
with the body will also differ with direction, and the body will be
optically anisotropic. Anisotropy markedly complicates the solutions of the equations of electromagnetism [7] because to any given
direction of propagation there correspond two different velocities of
propagation of the light. Such material is said to be "birefringent."
The electric displacement vectors for these two different modes of
propagation are perpendicular to each other and to the propagation
direction. Along certain directions, called "optic axes," the veloci-

350

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

ties of the two modes with perpendicular polarization are equal. A


material with only one such axis is said to be "uniaxial."
The components of the refractive index tensor can be complex,
resulting in anisotropic attenuation by absorption or scattering of
the incident light. This phenomenon is called "dichroism."
Birefringence can also occur in multicomponent media, even
when the components are themselves isotropic but have different
refractive indices. This type of birefringence, which is of geometric
origin, is called "form birefringence." It occurs in dilute polymer
solutions, unless care is taken to match the refractive indices of the
polymer and the solvent. It is particularly apparent in media that
consist of layers of optically different materials arranged in a
periodic fashion [7].
An example of an anisotropic medium is a sheet of crosslinked
rubber that is stretched in simple tension. The stretch direction is
the optic axis. The backbone bonds of the rubber molecules will be
oriented predominantly in the stretch direction. If, as shown
schematically in Figure 9-2, a beam of polarized light is passed
perpendicularly through the sheet, the interaction of the light with
the atoms of the rubber will be different if the direction of polarization lies along or is perpendicular to the stretch direction. Therefore the corresponding refractive indices will also differ. Whether
the index is greater in the stretch or in the transverse direction
depends on the detailed molecular structure of the rubber monomer
units.
Because, in this situation, there is no unique value of the refractive index, we cannot use Equation 9-1 to describe the propagation
of light in the medium. In the simple case of the stretched rubber,
Equation 9-1 must be replaced by
(9-4a)
(9-4b)
Equation 9-4 can be used to see what happens to two beams
whose polarization directions are along the x and y axes as they
pass through a sheet of thickness L. Initially, at time t = 0, at the
front face of the sheet, where z = 0, the amplitudes of the two

RHEO-OPTICS AND MOLECULAR ORIENTATION

351

Figure 9-2. Schematic diagram of polarized light passing through a sheet of rubber stretched
in a vertical direction; an example of an optically anisotropic material. (a) Plane of electric
vector is in stretch direction. (b) Plane of electric vector is in transverse direction.

electric vectors are given by


Ex = Ax Re[ exp i()x] = Ax cos ()x
Ey = Ay Re[exp i()y] = Ay cos ()y

352

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Therefore the two beams initially differ in phase by an amount


(8 x - 8 y )' After passage through the sheet, at z = L, the amplitudes are

The difference in phase angles is now

One beam has been "retarded" in phase by the amount


8

wL(ny - nJ/c
(9-5)

In Section 9.2 this result will be applied to the measurement of


birefringence.
An optical apparatus consists of a train of optical devices such as
polarizers, retarders, etc. An optically active test material is inserted into the apparatus as part of that train. In order to determine its optical properties from its response in the apparatus,
calculations of the sort that we have illustrated, leading to Equation
9-5, are made. Fortunately some elegant short cut methods have
been developed to do these otherwise tedious calculations. These
are the Mueller and the Jones matrix methods [6,7,11], in which
the state of the light is represented as a vector whose components
depend upon the components of the electric field vector. An optical
element is represented by a matrix. The effect of the interaction of
the element upon the incident beam is obtained by multiplying the
electric field vector by the element matrix and examining the
components of the resulting vector. The effect of a train of optical
elements is obtained by multiplication of the matrices of the elements. For a full description of how to use these methods the
reader is referred to the literature cited [6, 7, 11].

RHEO-OPTICS AND MOLECULAR ORIENTATION

353

9.2 MEASUREMENT OF BIREFRINGENCE

Birefringence is generally determined by measuring the intensity of


polarized light that is transmitted by an apparatus that includes the
sample as part of the optical train. We saw in Section 9.1.2 that
passage of polarized light through a birefringent medium causes a
relative phase shift (or retardation) of the two components of the
electric field vector of the light. Figure 9-3a shows an optical
arrangement in which light is first passed through a "polarizer," an
optical element which transmits only light that is linearly polarized
in a specific direction normal to the beam. The light then goes
through the test sample, which retards one component of the
electric vector, thus transforming the state of polarization of the
light from linear to elliptical. The emerging beam is then passed

(a)

(b)

Figure 9-3. Schematics of optical arrangements for the measurement of birefringence.


(a) Plane polarized light. (b) Circularly polarized light.

354

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

through an "analyzer." This is an element identical to the polarizer,


but with its transmission direction at right angles to that of the
polarizer.
The intensity of the beam exiting from the analyzer is proportional to the square of the amplitude of its electric vector. It is easy
to show, by use of the Jones or Mueller matrix methods [6,7,11],
that the ratio R of the transmitted intensity to the incident intensity
is given by
(9-6)
The angle I) is given by Equation 9-5, which involves the principal
components of the refractive index tensor nx and n y. f3 is the angle
between the polarization direction of the incident beam and the
faster principal direction of the refractive index tensor, and it is
also the angle of inclination of the polarization ellipse to the
laboratory coordinate axes; I) is proportional to the difference of
the lengths of the axes of the ellipse.
Let us now consider the implications of Equation 9-6. If a
birefringent material is placed between a polarizer and an analyzer,
and if both the polarizer and analyzer are rotated, while keeping
their polarization directions perpendicular, the transmission intensity will vary with the angle of rotation, as shown in Figure 9-4. The
transmission is zero when f3 = 0, 7T /2, 7T, etc.; these are the
extinction directions. The maximum transmission occurs at f3 =
7T / 4, etc. and is equal to sin 2( I) /2).
A more interesting situation is one in which the birefringence
and the extinction directions are not uniform, but vary over the
field of view. A good example of such a situation occurs in the flow
of a polymer melt into a sudden contraction [12]. Extinction patterns for this flow are shown in Figure 9-5. According to Equation
9-6 extinction occurs when either sin(I)/2) or sin(2f3) is zero. These
two possible causes for extinction can be distinguished in a number
of ways. Figure 9-3b shows a modification of the optical arrangement that incorporates "quarter wave plates," suitably oriented.
Quarter wave plates are birefringent elements that shift the relative
phase of the two components of an incident, linearly polarized
beam by 7T /2 radians, thereby converting it to a circularly polarized
beam. For this apparatus the transmitted intensity ratio is given by

RHEO-OPTICS AND MOLECULAR ORIENTATION

355

0.8

0.6

0.4

0.2

rI
/

//

/ ' .....\

~;';'

~/

---.......

\
\

\\

',~

\
rI
' I

______________________________
","
',~
'

O~~_;'

'
","

~,,'"

;'

w/2

\
\

...--...........,~

J'"____________

"'~~~~

w/4

/
/
/

/".'\

"

_______________
,~-...

3w/4

Figure 9-4. Relative transmitted intensity versus the angle f3 between the polarizer and the
principal direction of the refractive index tensor for various values of the retardation, 15, from
Equation 9-6. Continuous line, 15 = 7r. Long dashes, 15 = 7r /2. Short dashes, 15 = 7r / 4.

Equation 9-7:
(9-7)
In contrast to Equation 9-6 the orientation angle {3 no longer
affects the transmission, and R is thus a function only of the
retardation. If white light is used to illuminate the material the
transmitted light will consist of bands of different color, because
extinction will occur at any point only for light of the wavelength A
that satisfies both Equations 9-4 and 9-6. Bands of the same color
correspond to the same retardation, and are therefore called "iso-

Figure 9-5. Fringe patterns due to birefringence of HDPE at 2000C flowing through sharp
contraction. (a) Q = 3.8 cc/min. (b) Q = 17.3 cc/min. Reproduced from Ref. 12. Copyright
1973 by John Wiley & Sons, Inc. Reprinted by permission.
356

RHEO-OPTICS AND MOLECULAR ORIENTATION

357

chromatics." With the apparatus of Figure 9-2a, the extinction


bands that arise from the condition sin(2m = 0 will rotate if the
polarizer/analyzer combination is rotated, because {3 depends on
the angle of inclination of the polarizer to the principal axis of the
refractive index tensor. These bands are therefore called "isoclines."
Evidently, observation of the extinction bands by such an apparatus gives an immediate pictorial view of the birefringence pattern in
a complex flow field. However, by use of the "stress-optical relation,"
discussed further in Section 9.3, a quantitative analysis of the
birefringence leads to a measurement of the components of the
stress tensor in such a flow field. This is the principal application of
the flow birefringence measurement. The work of Han and Drexler
[12] gives a good example of such an application; Mackay and Boger
[2] give references to numerous other such studies.
It is clear that more than one measurement is required to
determine the two angles that characterize the refractive index
tensor ellipse for the two components in the plane normal to the
light direction. If the test specimen is stationary, as in our example
of a sheet of stretched rubber, this is easy to do experimentally. A
polymer solution or melt in a steady flow field can also be studied in
this manner.
It has not been possible, until recently, to measure the birefringence tensor under transient conditions, but a number of methods
applicable to transient measurements have now been developed.
The principles and optical arrangements of these are reviewed by
Fuller [1]. In general they include in the optical train an element
whose optical characteristics are varied rapidly compared to the
rate of the optical changes of the material being tested. One such
apparatus [13] uses a photoelastic modulator, a retarder whose
retardance is modulated at a frequency of 50 kHz. Another [14]
uses a polarizer or retarder with constant properties, but which is
rotated mechanically at 10 kHz. An application of the photoelastic
modulator apparatus to the transient response of concentrated
polymer solutions illustrates the utility of this apparatus [15].
We have only described the principles of the various measurement techniques. For detailed discussions of these principles, of the
apparatus and of sources and analyses of error, we refer the reader
to the literature cited.

358

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

9.3 BIREFRINGENCE AND STRESS

9.3.1 Stress-Optical Relation

The use of birefringence results to infer the components of the


stress tensor in a flowing or deformed polymer depends upon the
existence and validity of a so-called "stress-optical relation." This
relation is expressed as a proportionality between the components
of the refractive index and stress tensors:
(9-8)
The constant C is the "stress optical coefficient." The sign and
magnitude of C depend on the chemical structure of the polymer,
being governed by the "polarizabilities" of the bonds between the
atoms of the polymer molecule and by the directions of the bonds
with respect to the polymer backbone. The polarizability is a
measure of the strength of the response of the electrons in a bond
to an electric field. It depends on the number of electrons in the
bond and on the atoms being bonded. For a flexible polymer C is
proportional to the difference of the polarizabilities parallel and
transverse to the polymer backbone.
For a given polymer, C is essentially independent of the molecular weight and its distribution, and is relatively insensitive to temperature. Tabulations of experimental data on polymer melts are
given by Janeschitz-Kriegl [16-Section 1.2.3.2] and by Van Krevelen
[17].
The application of Equation 9-8 to a simple shear flow results
[16, p. 63] in the relations
=

(an/C) sin 2X

(9-9a)

T22 =

(an/C) cos 2X

(9-9b)

2T12
Tn -

where an is the birefringence and X is the extinction angle,


measured with the light directed along the neutral 3-direction,
with the fluid velocity in the 1-direction and its gradient in the
2-direction.

RHEO-OPTICS AND MOLECULAR ORIENTATION

359

In principle, the validity of the stress-optical relation should be


established experimentally for the particular polymer, for the kinematics of the deformation, and for the magnitude of the stresses
encountered. In practice, the rule has wide applicability to melts of
homogeneous polymers that are isotropic at rest. Heterophase
mixtures, including solutions, are excluded because there is a strong
contribution of form birefringence, which arises from the difference
in refractive indices of the components. Liquid crystalline polymers
are birefringent in the absence of stress and do not obey Equation
9-8 [18-21]. The only reported case of a homogeneous melt of a
flexible coil polymer failing to obey the relation appears to be that
of polystyrene in a high stress extensional flow, very close to the Tg
of the polymer [22].
The historical background and theoretical basis for the stressoptical relation are reviewed at length by laneschitz-Kriegl [16] in
his monograph on the subject. He also reviews the experimental
verification of its range of validity and illustrates its usefulness in
testing the predictions of various theories of viscoelasticity. The
theoretical basis for the relationship among orientation, stress, and
birefringence is also summarized by Doi and Edwards [23].
The stress-optical relation can be derived for a variety of molecular models. These all have in common the following elements:
1. The system is composed of optically anisotropic elements, for
example, segments of a polymer chain or dumbbells suspended
in a liquid. The optical contributions of the individual elements, expressed as polarizabilities, are additive.
2. The spatial distribution of orientations of the elements in the
absence of stress is Gaussian because of Brownian motion.
3. The stress in a deformed system is proportional to a tensor
that describes the orientation distribution of the elements.
The refractive index tensor is then proportional to the same
tensor, and it is this proportionality that leads to Equation 9-8.
It should be noted that the proportionality of stress and refractive index does not imply or require proportionality of stress and
strain rate. The stress-optic relation holds for melts well into the
non-Newtonian region of flow behavior [16]. The only requirement
is that both the stress and the refractive index are governed by the

360

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

orientation distribution. This will be true as long as the flow does


not change the orientation distribution, or the shape of a polymer
coil, too far from a Gaussian distribution. The failure of the
stress-optical relation observed in a high stress extensional flow [22]
is probably explained by the high chain extension expected under
these conditions. The inapplicability to liquid crystal polymers [18]
is due to the strong orientational ordering by intermolecular forces,
which overwhelms Brownian randomization; a similar result is
predicted theoretically for suspensions of rigid rods not subject to
Brownian motion [24].
The magnitude of the birefringence I1n is, as we have stated, a
measure of the deformation of a polymer coil. In other words, it
measures how much the distribution of chain segments varies from
being isotropic. The extinction angle X is determined by the direction of the axis of the deformed coil. For "weak" flows, such as
simple shear (see Section 5.8), this level of characterization of
deformation and orientation is probably adequate. It is not clear
that it is adequate for "strong" flows, which produce considerable
deformation. It is possible that there may be more than one
distribution of orientations with the same average end-to-end distance, and that these would not, therefore be distinguishable solely
on the basis of birefringence measurements.
An example from the dilute solution field may help to illustrate
the possibility that two different orientation distributions can lead
to the same birefringence. It is observed [25] that with an increasing
steady extension rate there is a region in which the birefringence
increases rapidly. This phenomenon is the "coil-stretch transition."
Two very different models for the deformation of a polymer coil in
a dilute solution have been proposed for this transition. These are
illustrated in Figure 9-6. In the "yo-yo" model [26] most of the
stretching occurs initially by orientation of the polymer chain segments around the middle of the chain, with the segments at the
ends remaining relatively unoriented. Eventually, at a very high
extension rate, the entire chain stretches out to its maximum extent.
Another model [27] suggests, however, that the stretching occurs
relatively uniformly along the chain. With increasing time the coil is
deformed in stages, shown schematically in Figure 9-6. Both models
are consistent with a sudden increase of birefringence. Other measurements, such as light scattering, would be needed to distinguish

RHEO-OPTICS AND MOLECULAR ORIENTATION

361

Figure 9-6. Models for process of the stretching of a polymer molecule by a high strain rate
elongational flow. (a) "yo-yo" model of Ryskin [26]. (b) Finitely extensible nonlinear spring
(FENE) bead chains of Wiest and Bird [27].

362

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

which model, if either, predicts more correctly the shape of the


deformed molecule.
9.3.2 Applications of Birefringence Measurements

Many of the applications of birefringence measurements have been


of theoretical interest, and are reviewed in the references
[1,2, 15, 16,23]. These measurements have been made to test the
predictions of various theories of viscoelasticity. They have also
been useful to measure rheological properties, such as the second
normal stress difference and the transient first normal stress difference, which are difficult to measure by mechanical means.
Visualization of the stresses in complex flows has been an important application of flow birefringence observations. The classic
studies of Tordella [28] on flow instabilities of polymer melts show
the usefulness of even qualitative observations to gain insight into
mechanisms, and to show differences in stress response arising from
differences in molecular structure. More recently, Baird has used
birefringence measurements to test the results of finite element
calculations for the stress field in an entrance flow [29].
Analysis of the stresses in injection molded parts by measurement of their birefringence permits inferences to be drawn about
the stresses and flow fields during molding [16, Ch. 4]. Allowance
has to be made for the thermal stresses during solidification of the
part. The existence of a maximum in tin near, but not at, the mold
wall has been attributed [30] to the occurrence of "fountain flow" in
the mold filling process.
The birefringence in an injection molded part can, as discussed,
be a useful indicator of the nature of the mold filling process.
Because it is a measure of "frozen-in" orientation it also correlates
with the tendency of a part to warp. In one application, however,
the birefringence is itself a property that must be minimized for the
application to be successful. That is in the use of optical data
storage disks. Understanding the causes of birefringence and of the
influence of processing conditions is vital, not merely useful, in this
case [31].
A final example is the somewhat surprising application of flow
birefringence measurements to the understanding of the melt spinning process. The surprising feature is that this process involves the

RHEO-OPTICS AND MOLECULAR ORIENTATION

363

highly transient, in stress and in temperature, draw-down of the


melt to the glass transition temperature. These are the types of
conditions for which the stress-optical relation failed in another
experiment [22]. Nevertheless, the birefringence measured in fibers
spun over a wide range of conditions was found to be directly
proportional to the stress at the freeze point; the constant of
proportionality agreed well with that estimated from independent
measurements of the stress-optical coefficient [32]. The birefringence also correlated well with the mechanical properties of the
fibers. The stress at the freeze point turned out also to be useful for
the prediction of crystallization in the spinline [33].
These applications are intended as illustrations of the use of
birefringence measurements in polymer processing. To list possible
applications to all of the processes by which polymer melts are
shaped is beyond the scope of this chapter. It is hoped that this
short list will serve as a guide and as a stimulus to readers.
REFERENCES
1. G. G. Fuller, "Optical Rheometry," to be published in Annual Reviews of
Fluid Mechanics (1990).
2. M. E. Mackay and D. V. Boger, "Flow Visualisation in Rheometry," Chapter
14 in Rheological Measurement, A. A. Collyer and D. W. Clegg, eds., Elsevier,
New York, 1988.
3. J. V. Lawler, S. J. Muller, R. A. Brown and R. C. Armstrong, J. Non-Newt. Fl.
Mech. 20:51 (1986).
4. R. J. Binnington, G. J. Troup and D. V. Boger, J. Non-Newt. Fl. Mech. 12:255
(1983).
5. G. G. Fuller, J. M. Rallison, R. L. Schmidt and L. G. Leal, J. Fluid Mech.
100:555 (1980).
6. W. A. Shurcliff, Polarized Light, Harvard University Press, Cambridge, Mass.,
1966.
7. A. Yariv and P. Yeh, Optical Waves in Crystals, Wiley-Interscience, New York,
1984.
8. H. C. van de. Hulst, Light Scattering by Small Particles, Dover, New York, 1981.
9. M. Kerker, The Scattering of Light and Other Electromagnetic Radiation,
Academic Press, New York, 1969.
10. A. Onuki and M. Doi, J. Chem. Phys. 85:1190 (1986).
11. R. M. A. Azzam and N. M. Bashara, Ellipsometry and Polarized Light, NorthHolland, Amsterdam, 1977.
12. C. D. Han and L. H. Drexler, J. Applied Polymer Sci. 17:2329 and 2369 (1973).
13. P. L. Frattini and G. G. Fuller, J. Rheology 28:61 (1984).

364

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

14. G. G. Fuller and K. J. Mikkelsen, 1. Rheology 33:761 (1989).


15. D. S. Pearson, A. D. Kiss, L. J. Fetters and M. Doi, 1. Rheol. 33:517 (1989).
16. H. Janeschitz-Kriegl, Polymer Melt Rheology and Flow Birefringence, SpringerVerlag, New York, 1983.
17. D. W. Van Krevelen, Properties of Polymers, Table 10.3, Elsevier, New York,
1976.
18. K. F. Wissbrun, 1. Rheol. 25:619 (1981).
19. G. Kiss and R. S. Porter, Mol. Cryst. Liq. Cryst. 60:267 (1980).
20. Y. Onogi, J. L. White and J. F. Fellers, 1. Non-Newt. Fl. Mech. 7:121 (1980).
21. T. Asada, H. Muramatsu, R. Watanabe and S. Onogi, Macromolecules 13:867
(1980).
22. T. Matsumoto and D. C. Bogue, 1. Polym. Sci., Polymer Phys. Ed. 15:1663
(1977).
23. M. Doi and S. F. Edwards, The Theory of Polymer Dynamics, Sections 4.7 and
7.2, Oxford University Press, Oxford, 1986.
24. L. G. Leal and E. J. Hinch, Rheol. Acta 11:190 (1972).
25. G. G. Fuller and L. G. Leal, 1. Non-Newt. Fl. Mech. 8:271 (1981).
26. G. Ryskin, 1. Fluid Mech. 178:423 (1987); Phys. Rev. Letters 59:2059 (1987).
27. J. M. Wiest and R. B. Bird, "On Coil-Stretch Transitions in Dilute Polymer
Solutions," Rheology Research Center Report RRC 116, University of
Wisconsin-Madison, May 1988.
28. J. P. Tordella, "Unstable Flow of Molten Polymers," Chapter 2 in Rheology,
Vol. V, F. R. Eirich, ed, Academic Press, New York, 1969.
29. S. A. White and D. G. Baird, 1. Non-Newt. Fl. Mech. 29:245 (1988).
30. M. R. Kamal and V. Tan, Polym. Eng. Sci. 19:558 (1979).
31. S. Anders, H. Schmid and K. Sommer, Kunststojfe 79: 55 (1989) (English text
on page 20 of Translation section of Gennan Plastics edition).
32. H. H. George, Polym. Eng. Sci. 22:292 (1982).
33. H. H. George, "Spinline Crystallization of Polyethylene Terephthalate," Chapter 10 of High Speed Spinning, A. Ziabicki and H. Kawai, eds., Wiley, New
York, 1985.

Chapter 10
Effects of Molecular Structure
10.1 INTRODUCTION AND QUALITATIVE OVERVIEW OF MOLECULAR THEORY

At various places in previous chapters we have discussed molecular


theories of polymer melt rheology and the relationship of rheological parameters to molecular structure characteristics such as molecular weight. This chapter presents, for convenience of reference,
some useful relationships in summary form. While there is some
duplication of material given earlier, it is felt that the reader will
find this useful. Numerical values of parameters are presented in
the text to illustrate principles, but extensive tables of data are not
given. References are given to literature in which data compilations
are available.
The rheological behavior of all flexible chain polymers is remarkably similar in general. This qualitative similarity results from the
fact that the chains are very long and flexible, and that the path of a
chain in motion cannot cross a neighboring chain. The values of the
rheological parameters of a given polymer depend on the details of
its molecular structure and molecular weight distribution.
The molecular theories discussed are based on three fundamental concepts:
1. Flexible polymer chains adopt the shape of a random coil in
the absence of a stress.
2. The coil is viscoelastic. It changes its shape under the influence of deforming stresses. When these stresses are released it
tends to return to its original shape) but the return is retarded
by molecular friction.
3. the density of the coils in the space that they occupy is much
smaller than the observed density. Therefore, the coils overlap
365

366

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

(a)

(b)
STRESS
h

RELEASE

Figure 10-1. Schematic illustrating concepts on which molecular theories are based: (a)
Rapid rearrangement of chain conformation by rotation about backbone bond by Brownian
motion results in loss of correlation of spatial relation of atoms 1 and 2, and leads to
formation of random coil. (b) When a deforming stress is released, Brownian motion restores
the random coil shape; restoring motion is retarded by molecular friction.

considerably, so that the motion of any individual molecule is


strongly affected by the presence of its neighbors.
These concepts are illustrated schematically in Figure 10-1 and are
discussed briefly in this section.
Most of the polymers of interest to us are composed of long
backbone chains of atoms such as carbon, oxygen, nitrogen, etc.
The bonds connecting these atoms are generally not linear but are
arranged at various characteristic angles. At sufficiently high temperatures the chain atoms are free to rotate rapidly, in a time on
the order of 10- 9 seconds, keeping the bond angles fixed. In a very
short time, therefore, a polymer chain will change the details of its
arrangement in space ("conformation") many times. Furthermore,

EFFECTS OF MOLECULAR STRUCTURE

367

even if we fix the position of one chain atom, we rapidly lose all
knowledge of the other atoms down the chain.
The spatial distribution of the chain atoms can be modeled
approximately by the statistics of a random walk in three dimensions. The direction of each "step" in such a walk is independent of
the direction of the preceding step. The result of this process is that
the chain adopts, on the average, the shape of a spherical coil. The
chain is almost never fully extended and almost never is it folded
into a compact arrangement such as in a crystal. On the average the
square of the distance between the chain ends is given by:

(10-1)
where <R2 > is the average of the square of the distance, N is the
number of steps, and I is the length of a step.
In practice the atoms are not completely free to rotate as we have
described. Their motion is influenced by the presence of other
atoms, either on the backbone chain or on groups of atoms attached to the backbone. The strength of the interactions with other
atoms depends upon the detailed molecular structure of the polymer and is responsible for the different magnitudes of rheological
and mechanical properties, such as viscosity and stiffness. In effect
the presence of strong interactions or of bulky pendant groups
attached to the backbone stiffens the chain and makes it less
flexible. In terms of the "random walk" model for the formation of
coils, a stiff chain takes a smaller number of steps than a more
flexible chain of the same total length, but each step is longer. The
product of the number of steps and the length per step is equal to
the total chain length in each case. Nevertheless, if a chain is
sufficiently long, that is, if it is long enough to contain a large
number of steps, it will adopt the random coil configuration.
Suppose that we stretch such a coil by applying an external force
so that more of the bonds lie along some preferred direction than
in the two orthogonal directions. Brownian motion tends to restore
the original random arrangement, and in effect sets up a restoring
force opposed to the applied force. If the external force is removed,
the coil returns to its undisturbed original shape. This randomizing
tendency of Brownian motion is the origin of the "rubbery"

368

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

elasticity of flexible polymers. The elastic recovery does not occur


instantaneously. It is slowed by the intermolecular forces that are
responsible for intermolecular friction and thus the viscosity of low
molecular weight fluids.
The mechanics of such a coil have not been worked out exactly.
The approximate model developed by Rouse and generalized by
Zimm and Bueche is described in Section 2.11.1. These bead/spring
models were developed to explain the behavior of polymers in
dilute solution, where there is little or no interaction of the individual polymer coils. In order to understand the rheology of polymer
melts we need to consider the effect of the strong interactions
between coils in concentrated systems.
The experimentally demonstrated disparity between the measured density of a molten polymer and the density calculated from
the measured volume occupied by a single polymer coil leads very
naturally to the qualitative but useful concept that the polymer
molecules interact by forming "entanglements." The topology of
entanglements, the reptating motion suggested by this topology, and
the results of the Doi-Edwards theory based on these concepts, are
described in Section 2.11.2.
To close this overview, it is worth emphasizing that the molecular
models discussed predict that the qualitative behavior of all flexible
linear polymer chains is the same. It arises from their conformation
as random coils that can overlap and interfere with each other's
motions if the molecular weight is sufficiently high. The quantitative
aspects of their rheology-the magnitude of the viscosity and the
elastic compliance, the critical molecular weight required for entanglement to occur, the temperature dependence of viscositydepend upon the detailed chemical structure of the segments of
which the chains are composed.
10.2 MOLECULAR WEIGHT DEPENDENCE OF ZERO SHEAR VISCOSITY

When the viscosity is measured at a sufficiently low shear rate the


viscosity approaches a constant value, denoted 7]0. In almost every
case studied 7]0 is independent of the molecular weight distribution
(MWD) and depends only upon the weight average molecular
weight (Mw). This dependence is shown schematically in Fig. 10-2.
Below some critical molecular weight Me the viscosity is approxi-

EFFECTS OF MOLECULAR STRUCTURE

369

Figure 10-2. Molecular weight dependence of the zero shear viscosity.

mately proportional to Mw. Above Me the dependence becomes


much steeper, with the viscosity varying with the 3.4-power of Mw.
This dependence has been observed experimentally for a wide
range of polymer structures.
In equation form,

(1O-2a)
(1O-2b)
The critical molecular weight Me varies widely for different
polymers. A few typical values [1] are shown in Table 10-1. Also
shown are two related quantities, Me and M;, discussed further

370

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Table 10-1. Critical Molecular Weights for Rheological


Properties of Selected Polymers
Polymer

Me

Me

M'e

Polyethylene
Polybutadiene
Polyisobutylene
Polyvinyl acetate
Poly (methylmethacrylate)
Polydimethylsiloxane
Polystyrene

1250

3800
5600
17000
22500

14400

6-10000

27500
29000
36000

> 150000

19000

130000

below. Me is the molecular weight between entanglements and M~


is the molecular weight above which the compliance Jl becomes
independent of molecular weight.
Mc appears to depend primarily on chain stiffness. This can be
seen from the values in Table 10-1, where the substitution of bulky
pendant groups in polystyrene and polymethyl methacrylate drastically increases Mc above that of polyethylene. Graessley and
Edwards [2] show a correlation with chain dimensions (expressed in
terms of the step length I in Equation 10-1), and also refer to other
correlations proposed earlier.
The other parameter needed to define the curve in Figure 10-2 is
the viscosity YJ cr at the critical molecular weight Mc' Van Krevelen
and Hoftyzer [3] give data for a number of polymers and discuss the
temperature dependence of viscosity. They also give an empirical
correlation of YJ cr with a quantity that is a function of the contributions from individual chemical groups (e.g., methylene, phenyl, etc.)
of which the polymer is composed. In addition to this group
contribution term, this correlation requires knowledge of the glass
transition temperature Tg of the polymer. When dealing with a
polymer of a new chemical structure this correlation is useful for
estimating the viscosity.
10.3 COMPLIANCE AND FIRST NORMAL STRESS DIFFERENCE

The compliance of a polymer melt, which is a measure of its


elasticity, was introduced in Section 2.7. The dependence of Jl on
molecular weight for a monodisperse polymer is sketched in Figure

EFFECTS OF MOLECULAR STRUCTURE

371

109 M

Figure 10-3. Molecular weight dependence of the steady state compliance for a monodisperse polymer.

10-3. At low molecular weights J1 is proportional to the molecular


weight M, but above a critical molecular weight M; it becomes
constant. The dependence of M; on molecular structure is similar
to that of Me' and is illustrated in Table 10-1. Below M; the
magnitude of J1 is approximately equal to the Rouse theory
prediction:

J1 =

OAM
pRT

for M < M;

(2-101)

Above M;, on the other hand, the compliance is approximately


constant, and is given by:

JO
S

oAM;

= ---

pRT

for M > Me'

(10-3)

where p is the density, R the gas constant, and T the absolute


(Kelvin) temperature. Above M; one may think of the melt as

372

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

though it were a rubber crosslinked by entanglements, with a


molecular weight between crosslinks determined by M~.
The constant value J1(M~) at high molecular weights is also
related to the plateau modulus Gt of the polymer that is determined by stress relaxation or by small amplitude oscillatory shear
experiments. The product J1Gt is approximately constant at a
value of about 3. This constancy follows naturally if one recalls that
Gt is expressed by Equation 2-20, similar in form to the reciprocal
of Equation 10-3.
GO

pRT

=--

Me

(2-20)

The constancy of the product J1Gt implies a close relationship


between Me and M;, as shown in Table 1O-l.
The compliance J1 and other manifestations of melt elasticity,
such as normal stress difference, extrudate swell, capillary flow end
correction, and also, as we will see below, shear rate dependence of
viscosity, are extraordinarily sensitive to molecular weight distribution (MWD). This was discovered very early in the study of polymer
melt viscoelasticity [4] and is illustrated in Figure 10-4. Two polymers of different molecular weight but approximately equal compliances are blended; the blends have a compliance that is as much as
ten times that of either component.
The extreme sensitivity to the MWD, especially to small amounts
of very high molecular weight components, has made exact determination of the relationship between J1 and MWD difficult. The best
estimates fit the form shown by Equation 10-4

(l0-4)
Here, J1(M;) is the value from Equation 10-3. The factor f is a
function of the MWD. Among the forms for this factor that have
been proposed are:

f=

(10-5)

EFFECTS OF MOLECULAR STRUCTURE

373

100~------------------------------~

40 f20 f- 0

o
o
o

c
I

0.2

0.4

0.6

0.8

1.0

WEIGHT FRACTION B

Figure 10-4. Compliance of blends of two silicones: M,)A = 5.85 X 10 4 , Mw)B = 5.96 X 10 5.
Adapted from Ref. 1. Copyright 1984 by The American Chemical Society. Reprinted by
permission.

and
2 < a < 3.7

(10-6)

The Curtiss-Bird model prediction given by Equation 2-115 involves


even higher molecular weight averages than these expressions.
The high molecular weight components of the MWD make the
greatest contributions to both of these expressions for f, and they
are also the most difficult to detect precisely. Furthermore, for
some MWDs the expressions are indistinguishable. For example,
for the so-called "log normal MWD" (see Appendix B) f reduces
to (Mw/Mn)3 if calculated from either Equation 10-5 or 10-6, if a is
equal to 3. Blending experiments also can not distinguish between
the two above forms [5]. In any case, it is clear that IJ is extremely
sensitive to MWD, and especially to the fraction of very high
molecular weight polymer. Measurement of IJ, or of some associated viscoelastic property, is therefore a far more sensitive method

374

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

for characterizing polymers for subtle differences in MWD than are


dilute solution measurements of MWD.
According to Equations 10-3 and 10-4, J~ is proportional to the
absolute temperature. Therefore, for the normal processing range
of polymer melts, J~ does not vary much with temperature, especially as compared to viscosity.
The steady state compliance is a measure of elasticity in the
linear range of viscoelastic behavior. As shown by Equation 3-42, in
the limit of very low shear rates it is related to the first normal
stress coefficient '1'1,0'
(3-42)

Recalling the definition of '1'1 (Equation 4-6), in this limit Equation 3-42 can also be written as a relationship between the first
normal stress difference N1 and the shear stress u.
N 1 -- 2J".2
sv

(10-7)

In the nonlinear range the ratio N1/2u 2 is no longer the steady


state compliance. To a first approximation this ratio can be regarded as independent of shear rate.
10.4 SHEAR RATE DEPENDENCE OF VISCOSITY

Typically the viscosity of molten polymers approaches a constant


value 710 at low shear rates, and begins to deviate from 710 at some
characteristic shear rate Yo' At higher shear rates the viscosity
approaches "power law" behavior, i.e., a straight line with negative
slope on a log-log plot. The~y.ariation of viscosity with shea!_Jat~ is
very dependent upon the molecular weight distribution ~(MWD).
The magnitude of the characteristic shear rate, the power law
slope, and the detailed shape of the viscosity-shear rate curve all
depend upon the MWD.
The qualitative relationship of the shapes of the flow curves and
of the MWD curves is illustrated in Figure 10-5 [6]. Qualitatively
the flow and MWD curves are mirror images of each other. The
data for this illustration were obtained with concentrated polymer
solutions, but similar relationships have been found for polymer

EFFECTS OF MOLECULAR STRUCTURE

375

1.0.---""7------====--.,

(a)

i=

()

<t:

IE

I-

I
I
I

0.5

:r:

(!)

Lij

:!:

.... '"
Ow.~
....______
o
5 x

_L_ _~----~~--------~~~

10'

1.5 x 10 5

10 5

MOLECULAR WEIGHT

(b)

1.0

---~------

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

",

';;: 0.5

\
\

O~

10

______

______

10 2

10 3

______

10'

____

10 5

SHEAR STRESS, Pa

Figure 10-5. Effect of molecular weight distribution on the shape of the viscosity curve.
Adapted from Ref. 6. (a) Integral molecular weight distributions of two polyisobutene
solutions; weight fraction of polymer below molecular weight M versus M. Solid line, broad
molecular weight distribution polymer; dashed line, narrow MWD fraction having a similar
value of Mw- (b) Viscosity curves for solutions of the two polymers whose MWDs are shown
in Figure 1O-5a. Volume fraction = 0.45.

376

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

105r-----------------------------------------~

102~

________~__________~________~________~
10
10 2
SHEAR RATE (5- 1)

Figure 10-6. Flow curves of three whole polymers, A, Band C, having most probable
molecular weight distributions, and that of a blend, D, of A and C. Broadening the
distribution by blending reduces the shear rate at which shear thinning begins. At high shear
rates, the viscosity of D approaches that of the lowest molecular weight whole polymer.
Adapted from Ref. 8. Copyright 1965 by Huethig & Wepf. Reprinted by permission.

melts [7]. The effect of MWD upon the shear rate for the onset of
shear thinning is demonstrated clearly by the results of a blending
experiment [8], as shown in Figure 10-6. Curves A, B, and C in this
figure are the flow curves of whole polymers with similar MWDs,
with an Mw/Mn ratio of about 2. Curve D is the flow curve of a
blend of polymers A and C; polymer D has an M w equal to that of
B. In accord with Equation 10-2, Band D have equal "70 values.
However, the viscosity of D begins to decrease with shear rate
much sooner than that of B. At high shear rates the flow curve
begins to approach that of the low molecular weight polymer C.
We now consider quantitative relationships for the dependence
of viscosity upon shear rate, discussed previously in Section 4.5.1.
The simplest such relationship is the power law equation 4-8:
(4-8)

EFFECTS OF MOLECULAR STRUCTURE

377

As discussed in Section 4.5.1, this equation neither describes the


viscosity-shear rate relationship accurately over a wide shear rate
range nor offers any significant advantage for numerical simulation.
More realistic expressions, such as Equations 4-11, 4-15 or 4-16,
were given in Chapter 4.

71(Y)

710[(1 + IAylm)]-l

(4-11)

71(Y)

710[1 + (Ay)2r p

(4-15)
(4-16)

All of these equations have similar shapes, approaching 710 at low


shear rates and power law behavior at high shear rates. The shear
rate region around y = 1/A is the appreciably curved region of the
pronounced onset of shear thinning. We can therefore discuss quite
generally, for all equations of this type, how the characteristic time
A and the power law slope depend on molecular structure.
Experimentally, Graessley [9] has shown for a variety of narrow
MWD linear polymers that there is an apparently universal relationship involving Y'Jo, JJ, and Yo, the shear rate for the onset of
shear rate dependence (defined here as the shear rate at which
71 = 0.8710)' as follows:

JJ71oYo

0.6( 0.2)

(10-8)

It has also been suggested that this correlation is applicable to


polydisperse polymers. According to the Rouse theory, the product
JJ710 is approximately 0.6 times the longest relaxation time AR , as
seen from Equations 2-98 and 2-101. We can use AR , therefore, to
estimate the approximate magnitude of A in Equations 4-11, 4-15,
and 4-16. From the discussion in Section 10.3, especially Equations
10-4 to 10-6, we recall that JJ, and therefore also AR , increase very
rapidly as the MWD is broadened. Equation 10-8 is consistent with
the general observation that polymers with broad MWDs show the
onset of shear thinning at low shear rates, as illustrated in Figures
10-5 and 10-6.

378

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Theory offers only limited guidance for the prediction of the


shapes of flow curves. Bueche and Harding [10] had proposed
Equation 4-13, a special case, with an exponent m of 0.75, of the
more general Equation 4-11 suggested later by Cross [11].
(4-13)
Their theoretical derivation suggested that A is equal to twice the
longest Rouse relaxation time AR' given by Equation 2-98.
(2-98)
However, it is not clear from experiment [8] precisely which
average should be used for the molecular weight M. Nor is an
exponent of 0.75 appropriate for all polymer melts. The Bueche
and Harding model emphasizes the importance of the dimensionless group Ai' for characterizing the shear rate dependence of
viscosity. It also suggests that the Rouse time AR is the appropriate
time constant.
A quantitative prediction of shear rate dependence of viscosity is
made in an early model by Graessley [12]. This model is based on
the idea that shearing decreases the number and effectiveness of
entanglements in the melt. This model also uses the longest Rouse
relaxation time as the characteristic time constant. Its predictions
have been confirmed experimentally for a number of systems, and
despite the fact that the theory behind it is no longer fashionable, it
is still a useful guide. It is particularly interesting to note that
Graessley found that Equation 4-13 is indistinguishable from his
theoretical prediction for polymers with a "most probable MWD."
These include condensation polymers, such as nylon, polyester, and
polyacetal, for which Equation 4-13 fits experimental data [8].
Graessley has also calculated the shear rate dependence for the
log-normal MWD, and his results have been fitted to Equation 4-11
[13]. The results of this curve fitting, over a range of Mw/Mn
between 1.05 and 12.6, are shown graphically in Figure 10-7 and
numerically in Table 10-2. In Reference 13 the curve fitting was
done with "reduced" variables 77 /770 and ARi'. In unreduced form

EFFECTS OF MOLECULAR STRUCTURE

379

10' r-------------------------~

i<-e-SLOPE

10 '

0.25

-~

A-'
f

SLOPE = -3

1.0
m

~\

10

10

10

4 L -______-.1________....1.______.........

0.1

\
10

100

Figure 10-7. Parameters of Equation (10-9) fitted to flow curves predicted by Graessley's
entanglement model for polymers with log-normal molecular weight distribution. Adapted
from Ref. 13. Copyright 1986 by The Society of Rheology. Reprinted by permi~sion of
John Wiley & Sons, Inc.

Table 10-2. Parameters of the Cross Equation (4-11) for Several


Log Normal Molecular Weight Distributions [13].
(Data simulated by means of Graessley's theory [12])

f3

(See Appendix B)

Mw/Mn

A -1
f

0.3
1.0
1.5
2.0
2.5

1.05
1.65
3.1
7.4
12.6

1.6
0.48
6.4 x 10- 2
3.0 x 10- 3
4.4 x 10- 4

0.842
0.704
0.564
0.456
0.423

380

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the result is:


(10-9)

where fl(M;) is the Rouse theory prediction for the compliance of


a monodisperse polymer with sufficiently high molecular weight, as
given by Equation 10-3.
OAM;

fO = - S

pRT

for M > Me'

(10-3)

The product 770fl is thus the time constant for a monodisperse


polymer with the same 770 as the polydisperse polymer for which
the theoretically predicted flow curve was fitted. The fitting parameter AI plotted in Figure 10-7 is, therefore, a quantitative measure
of how the breadth of the MWD shifts the flow curve along the
shear rate axis, compared to that of a monodisperse polymer.
The results of this curve fitting are most interesting. The slope m
decreases with increasing breadth of MWD (Mw/Mn), but the
dependence is not very sensitive at the higher values of Mw/Mn'
The slope m is therefore not a sensitive measure of the breadth of
the MWD. This means also that the shape of the flow curve is not
strongly affected by small changes in the MWD.
The time constant multiplier AI' on the other hand, is extremely
sensitive to MWD, varying as (Mw/Mn)3. We have identified the
characteristic time A with A R , and the product AIfl(M;) can
therefore be interpreted as the compliance f1 of the polydisperse
polymer. The third-power dependence of AI on Mw/Mn for the
log-normal MWD is consistent with this interpretation. We recall
from Section 10.3 that for this MWD fl also has this dependence.
With this interpretation AI is the factor by which fl(M~) must be
multiplied to obtain the actual fl,
( 10-10)

which is identical to Equation 10-4, identifying AI with !(MWD).


Equation 10-9 represents a "master curve," applicable in principle to any sample of a high molecular weight polymer of a given

EFFECTS OF MOLECULAR STRUCTURE

381

type. All of the molecular weight, temperature (and possibly pressure) dependencies are incorporated through 77o. The MWD dependence is expressed by Af and m. The effect of the chemical
structure of the polymer comes in through the dependence of 77o
and Jl(M;), discussed above.
A master curve such as Equation 10-9 is particularly useful for
extending the range of shear rates beyond that for which data can
be obtained. Typically a given rheometer is capable of covering only
a limited range of shear rates. By making use of the time-temperature superposition principle (Section 2.12), one can convert viscosity
versus shear rate data measured at various temperatures to a single
curve for a wider shear rate range at a fixed reference temperature.
It is also useful if one is dealing with a family of polymers with
varying molecular weights but a constant MWD. Condensation
polymers such as polyesters are a good example of such a family.
The product AfJsO(M;) is then a constant, independent of molecular weight and virtually independent of temperature. One can then
construct a master curve with only the temperature dependent zero
shear rate viscosity 770(T) as a parameter.
It should be noted that the particular values of Af and m shown
in Figure 10-7 and Table 10-2 apply in principle only to polymers
with a log-normal MWD. For blends of two such polymers the
appropriate value of Af should probably be estimated from Equation 10-5 or 10-6, using the blending rules given in Appendix B. The
dependence of m on MWD is sufficiently small for the data in
Figure 10-7 to give a reasonable estimate.
10.5 TEMPERATURE AND PRESSURE DEPENDENCE
10.5.1 Temperature Dependence of Viscosity

In order for a polymer melt to flow, the chain segments must be


able to move. This requires two conditions:
1. There must be space available for the motion to occur, and
2. The chain segments must have sufficient thermal energy to
overcome energy barriers that impede the motion, such as the
barrier to rotation about a valence bond.

382

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Every polymer has a characteristic temperature, the "glass transition temperature" Tg , which determines the first condition. A brief
discussion of the dependence of Tg on molecular structure is given
in Appendix D. In the vicinity of the Tg the availability of space,
i.e., free volume vf' is the limiting factor; at higher temperatures,
where there is no lack of free volume, the energy barriers become
significant.
Both experimentally and theoretically it is found [14] that near Tg
the dependence of viscosity upon free volume is described by the
Doolittle equation:
(10-11)

It has been suggested that vf increases approximately linearly


with temperature above Tg :
(10-12)

Here (Xf is the expansion coefficient of the free volume, approximately ((XL - (Xc), where (XL and (Xc are the thermal expansion
coefficients above and below Tg Substituting vf from Equation
10-12 into Equation 10-11, we arrive at the well-known WLF
(Williams-Landel-Ferry) equation [14]:

(10-13)

In the WLF equation, fo is the fractional free volume vf/v o at the


Tg of the polymer.
The WLF equation describes the temperature dependence of
viscosity well from Tg to about Tg + 100. Values of the parameters
of the equation, expressed in slightly different form, are tabulated
by Ferry [14, p. 316]. Ferry describes a graphical method for
determining the WLF parameters from experimental data.
It is most helpful to note that Equation 10-13 can be expressed in
a number of "universal" forms approximately applicable to all

EFFECTS OF MOLECULAR STRUCTURE

383

polymers. The most useful, although least accurate, form is Equation 2-129 with the parameters taken to be universal constants:
log TJ(T)

17.44(T - T)

log TJ(Tg) - 51.6 + T _ ~g

(10-14)

It is not uncommon to express the viscosity in the familiar


Arrhenius form, even though the activation energy Ea is not a
constant in the region of applicability of the WLF Equation,
TJ = A exp(Ea/RT )

(10-15)

The activation energy Ea is then given by


(10-16)
From Equation 10-16 we see that Ea depends both on the
absolute temperature T and upon the difference T - Tg. The
activation energy increases, i.e., the viscosity becomes more temperature dependent, as one approaches Tg . This is illustrated in Figure
10-8, in which Ea calculated from Equation 10-16 is plotted for an
assumed Tg of 373K (looOe).
Equation 10-16 can also be used to estimate the error in the
temperature dependence resulting from the use of the "universal"
form of the WLF Equation 10-14, instead of Equation 10-13, with
experimentally measured parameters. For example, for polystyrene
at 200 e the Ea from equation 10-13 is 40 kcal/mol K. From
Equation 10-16 the calculated value is 31 kcal/mol K.
At temperatures appreciably higher than Tg + 100, the temperature dependence of viscosity is no longer affected as strongly by the
increase of free volume. Instead, the energy barriers to motion
become limiting. In that case the temperature dependence is given
by Equation 10-15, with a constant Ea whose value depends upon
the chemical structure of the polymer.
As was mentioned in Section 10.2, Van Krevelen and Hoftyzer [3]
have proposed a more general viscosity-temperature relation and
have tabulated data for a number of polymers. However, it is worth
0

384

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

140r-----------------------------------,
e 120
(5

::::::"100
!'II

>a::

(!)

80

ffi

60

40

z
o

i=

20
O~~

:120

__

__

160

__

__

200

~~

__

240

__

__

280

~~

320

TEMPERATURE (OC)

Figure 10-8. Arrhenius activation energy versus temperature for WLF temperature dependence of viscosity, calculated from Equation 10-18 with Tg taken to be 100C (373 K).

noting that the WLF Equation 10-16 can give a reasonable estimate
of Ea at high temperatures. For example, consider polyethylene
terephthalate (PET), with a Tg of 70C. At 285C Equation 10-16
gives a value for Ea of 18 kcal/mol K. The experimental value is
13.5 [15], and Van Krevelen's calculated value [3] is 12 kcal/mol K.
Similarly, for acetal copolymer the calculated value from the WLF
Equation (assuming a Tg of - 60C) is 9.7, compared to a measured
value of 7.1 kcal/mol K at 190C. [8]. It appears that WLF overestimates E a , but gives at least a reasonable first approximation.
In order to give a feel for the effect of the magnitude of Ea on
the temperature dependence of viscosity, Figure 10-9 is a plot of
the percentage increase of viscosity caused by a 1C and a 10C
decrease of temperature at a melt temperature of 200C.
10.5.2 Pressure Dependence of Viscosity

We have just seen the close connection between Tg and the temperature dependence of viscosity. In Appendix D it is noted that Tg
depends on pressure. It is not surprising, then, that viscosity should

EFFECTS OF MOLECULAR STRUCTURE

385

1000r-----------------------------------~

>-

I-

ii5

o()

en

:; 100
~

(!)

Z
<I:
J:

()

~ 10
w

()

a:
w
Q.

1~~--~------~----~~----~----~
o
20
.40
60
80
100
ACTIVATION ENERGY (kcal)

Figure 10-9. Changes in viscosity (percent) due to temperature changes of 1C Oower curve)
and 10C (upper curve) as functions of activation energy at a melt temperature of 200C.

also be pressure dependent. There have not been many reliable


measurements of the coefficient of pressure dependence of viscosity
a, which is a difficult quantity to measure. An average value for a
from the literature is estimated to be on the order of 2 X 10- 8
m 2 jN (1.4 X 10- 4 psi-I), where a is defined by
1 d1]

a=--

1] dp

(1O-17a)

or, equivalently, by

1](p)

1](O)exp(ap)

(10-17b)

With this average value of a, a pressure of 10 4 psi would


increase the viscosity by a factor of four over the atmospheric
pressure value. This is roughly in agreement with the experimental
observations of Westover [16].
It should be noted that in the vicinity of Tg the effect of pressure
is much larger, just as is the effect of temperature near Tg Simi-

386

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

larly, near the crystalline melting point, pressure induced crystallization can cause solidification of a melt at a temperature above
the atmospheric pressure value of Tm'
Cogswell [17] expresses the pressure dependence of viscosity in
terms of an equivalent temperature change. An estimated value
for this pressure-temperature coefficient (dT jdp)lJ is -5 x 10- 7
m 2 KjN (-5 x 10- 8 cm 2 Kjdyne; -3.5 x 1O- 3Cjpsi). In other
words, applying a pressure of 1000 psi is equivalent to decreasing
the temperature by 3.5C. The exact value of this coefficient depends on the polymer.

10.6 EFFECTS OF LONG CHAIN BRANCHING (LCB)

Many polymers contain branches whose lengths are comparable to


the critical entanglement length of a linear polymer chain. Such
branches have a strong effect on various aspects of the polymer
rheology. The magnitude and even the direction of the manifestations of branching depend on the length of the branches, their
geometrical arrangement along the chain, and on their distribution
among the chains.
Sometimes branches are introduced deliberately by copolymerization with a multifunctional monomer. In many cases, however,
they occur as a result of side reactions during polymerization. In
these cases small variations of the polymerization conditions, catalysts, impurities, etc. may cause significant changes of the branch
concentration and therefore of the rheology.
The occurrence of branching is often accompanied by a broadening of the molecular weight distribution. The effects of branching
and of MWD variation may therefore be difficult to separate [18].
The most definitive information on the effects of branching has
been obtained from research on monodisperse polymers synthesized
with a known branch structure. Much of this work is reviewed by
Graessley [19], who has, with his collaborators, been a principal
contributor to this field.
Branching affects the zero shear viscosity YJo and the steady state
compliance 11- It also has an effect on the temperature dependence
of viscosity (E), on the extensional viscosity, and can be responsible for a dependence of the rheology on shear history. Because of

EFFECTS OF MOLECULAR STRUCTURE

387

the variability of the branching effects with the details of branch


structure, concentration, and distribution, it does not seem useful
to give more than a qualitative account here. Quantitative results
for specific polymers can be found in the references, but it should
be remembered that these are most often well determined only on
monodisperse model compounds.
Physically, the introduction of long branches has two opposing
effects. First, the radius of gyration R g is decreased compared to
that of a linear chain of the same molecular weight. This occurs
because the branch points tie together chain segments that might
be widely separated in space in a random coil of a linear polymer.
The decreased Rg results in fewer entanglements and a lower
viscosity.
The second effect of branching occurs when the branch length is
sufficiently long to be entangled, i.e., when the molecular weight of
the branch becomes comparable to Me the critical molecular weight
for entanglement of a linear chain. The overall entanglement network then has a much longer lifetime than that of a linear polymer
network. The viscoelastic relaxation spectrum is extended to much
longer relaxation times. This behavior is consistent with the reptation model. A linear chain relaxes by diffusing out of its entanglement "tube." A branched chain is attached at its branch points to
at least two other chains, each in their own tubes. It cannot
therefore diffuse out independently, and relaxation can occur only
by processes requiring much longer times.
At low molecular weights, and therefore also low branch lengths,
the zero shear viscosity 110 of a branched polymer is considerably
smaller, by as much as a factor of 10, than that of a linear polymer
of the same Mw' This is the result of the smaller coil radius. The
viscosity reduction is much less evident if the linear and branched
polymers are compared at the same intrinsic viscosity, because the
latter is also reduced by branching.
The enhanced entanglement efficiency of branches longer than
the critical entanglement length causes a large increase of 110 at
high molecular weights and branch lengths. The viscosity may be
orders of magnitude higher than that of a linear polymer at the
same Mw' The slope of the logarithmic plot of 110 versus Mw is also
increased from its value of 3.4 for linear polymers to double that or
more.

388

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The steady state compliance lJ is similarly affected by branching.


At low molecular weights it is lower than that of a linear polymer.
However, at high molecular weights it is enhanced and increases
above that of a linear polymer. Also, whereas for linear polymers
the compliance reaches a constant value lJ(M;) above a critical
molecular weight, that of branched polymers continues to increase.
If it levels off at all it does so at a much higher molecular weight
than the M; of the linear polymer and at a higher value of lJ.
The shear rate dependence of viscosity is affected by branching
because of its relationship to the compliance. Equation 10-10 appears to hold for branched as well as linear polymers. Therefore,
because of its higher lJ, the viscosity of a high molecular weight
branched polymer begins to decrease with shear rate at a lower
shear rate than that of a linear polymer of the same 710. It should
be remembered once more that for commercial polymers, which are
almost without exception polydisperse, branching is generally accompanied by broadening of the MWD, which also increases lJ
and shear sensitivity.
There are other effects of branching that are not yet well understood. Long chain branched polymers appear to have a higher
activation energy for viscosity than linear polymers of the same
chemical structure. This effect is most pronounced for polyethylene,
for which Ea is increased from about 6.5 kcaljmol K for the linear
polymer to about 14 kcaljmol K for branched polymers. Also, the
activation energy is found to depend on shear stress, rather than
being a constant.
Extensional viscosity is discussed in Chapter 6. However, it is
worth noting here that branched polymers such as LDPE show the
greatest degree of extension thickening. This has significant implications for processing, because of the effect on the stability of free
surface flows and the effect on the rate of stress-induced crystallization [20].
Finally, it has been observed in a number of cases that the
rheology of branched polymers is affected by shear history in a way
that is qualitatively different from that of linear polymers of the
same chemical type [21-23]. Shearing causes a decrease of the
apparent viscosity and the elasticity. The effect of shear is only very
slowly reversible with time. The effects are completely reversed by
dissolving and reprecipitating the sheared polymer. A possible

EFFECTS OF MOLECULAR STRUCTURE

389

explanation is that the branch entanglements that are responsible


for enhanced viscosity and elasticity are destroyed by high shear.
And just as these entanglements are slow to disengage by diffusion,
they can also take a very long time to reform once they have been
destroyed. The shear history effect is probably closely related to the
"shear refining" process for improving the performance of such
polymers as LDPE in processes such as film blowing.
REFERENCES
1. W. W. Graessley, "Viscoelasticity and Flow in Polymer Melts and Concentrated Solutions," Figure 24, Chapter 3 in Physical Properties of Polymers
J. E. Mark, ed., American Chemical Society, Washington, D.C. (1984).
2. w. W. Graessley and S. F. Edwards, Polymer 22:1329 (1981).
3. D. W. Van Krevelen and P. J. Hoftyzer, Properties of Polymers, Second Edition,
pp. 341-347, Elsevier, New York, 1976.
4. H. Leaderman, R. G. Smith, and L. C. Williams, 1. Polym. Sci. 36:233 (1959).
5. K. F. Wissbrun, Trans. Soc. Rheol., 21:149 (1977).
6. R. S. Porter, M. J. R. Cantow, and J. F. Johnson, Proc. Fourth Int'l. Congo
Rheol., E. H. Lee, ed., Vol. 2, 479 (1963).
7. H. L. Wagner and K. F. Wissbrun, SPE Trans., July, 222 (1962).
8. H. L. Wagner and K. F. Wissbrun, Makromol. Chemie 81:14 (1965).
9. W. W. Graessley, Adv. Polym. Sci. 16:1 (1974).
10. F. Bueche and S. W. Harding, 1. Polym. Sci. 32:177 (1958).
11. M. M. Cross, 1. Appl. Polym. Sci. 13:765 (1969).
12. W. W. Graessley, 1. Chem. Phys. 47:1942 (1967). See also Figure 28 of
Reference 1.
13. K. F. Wissbrun, 1. Rheol. 30:1143 (1986).
14. J. D. Ferry, Viscoelastic Properties of Polymers, Third Edition, John Wiley &
Sons, New York, 1980.
15. D. R. Gregory, 1. Appl. Polym. Sci. 16:1479 (1972).
16. R. F. Westover, SPE Trans. 1:14 (1961).
17. F. N. Cogswell, Plastics and Polymers, February, p. 39 (1973).
18. W. Minoshima and J. L. White, 1. Non-Newt. Fl. Mech. 19:251 (1984).
19. W. W. Graessley, Accounts of Chemical Research 10,332 (1977).
20. G. Perez, Paper presented at International Workshop on Extensional Flows,
Mulhouse, January 1983; G. Perez and C. Lecluse, Paper presented at 18th
Int'l. Man-Made Fibre Conf., Dornbirn, Austria, June 20-22, 1979.
21. J. H. Prichard and K. F. Wissbrun, 1. Appl. Polym. Sci. 13: 233 (1969).
22. M. Rokudai and T. Fujiki, 1. Appl. Polym. Sci. 23:3295 (1979).
23. M. Rokudai, 1. Appl. Polym. Sci. 26:1427 (1981).

Chapter 11
Rheology of Multiphase
Systems
11.1 INTRODUCTION

This chapter gives a brief account of how melt rheology is affected


by the presence of more than one discrete phase. Polymers filled
with rigid reinforcing agents such as fibers or minerals are such
systems, as are immiscible blends of polymers. Phase-separated
block or graft copolymers can be considered as immiscible blends,
with the added constraint that the blend components are joined
chemically. Foams are filled polymers in which the filler is a gas.
The relative briefness of this account is not because of the
simplicity of the subject or because there are not many phenomena
that are different from those observed with homogeneous polymer
melts. On the contrary, it is because the subject is so diverse and so
many effects are specific to the system components that we have not
attempted to do more than summarize the principal observations.
More detailed accounts can be found in books and review articles
devoted to this subject [1-7].
11.2 EFFECT OF RIGID FILLERS

The logical starting point for a discussion of the effects of rigid


fillers is Einstein's equation for the viscosity of a very dilute
suspension of rigid spheres in a Newtonian fluid. Einstein showed
that the presence of a sphere perturbs the flow field of the surrounding fluid in a shear field in such a way that more energy is
dissipated. The effect is to increase the viscosity by an amount
390

RHEOLOGY OF MULTI PHASE SYSTEMS

391

proportional to the volume fraction of the filler particles regardless


of their size:
(11-1)

Here 11 is the viscosity of the suspension, 11/ that of the suspending


fluid, and cf> the volume fraction of the filler.
Difficulties arise as soon as any of Einstein's assumptions no
longer apply. If the suspension is not very dilute, the flow field and
the particles interact with each other to further increase the suspension viscosity. If the particles are not spherical, the suspension is
no longer Newtonian and cannot be characterized by a single
viscosity. If the filler is not rigid, it is deformed by the flow, and the
suspension becomes non-Newtonian and viscoelastic. And of course
if the fluid medium is already a non-Newtonian and viscoelastic
material, such as a molten polymer, additional complications are
encountered.
There are other complexities of both theoretical and experimental origins. Density differences between the phases give rise to
buoyancy effects. In complex flows particle migration can occur, for
example due to the radial variation of shear rate in pipe flow.
Particles greater than 10 microns in diameter have little tendency to
aggregate, and their effect on the suspension is entirely a hydrodynamic one. However, if they are in a size range where colloidal
forces are significant, on the order of one micron or less, they will
be subject to strong attractive forces that tend to promote aggregation. Furthermore, the state of aggregation is altered significantly
by the presence of a deforming stress.
Experimental problems arise whenever the size scale of the
disperse phase is comparable to that of the measuring or processing
equipment. Highly concentrated dispersions pose particular difficulties in maintaining uniformity. For example, if the viscosity of
the suspending fluid is not sufficiently high, it can be squeezed out
of a rigid filled suspension when it is forced through a contraction,
leaving behind a cake or mat of filler that then acts as a filter.
Cheng [1] has reviewed some of these problems and their effects on
the precision with which the properties of concentrated dispersions
can be measured.

392

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

11 .2.1 Viscosity

Typically a fiber or mineral-filled polymer contains 30 weight % of


filler, corresponding to a volume fraction of about 20%. This
concentration far exceeds the range of applicability of the Einstein
equation 01-1) for the viscosity of a suspension. In such a concentrated suspension the disturbances of the flow field caused by the
filler particles interact with each other to increase the viscosity
much more than is predicted by Equation 11-1.
A large number of empirical and theoretical equations have been
proposed to describe the effect of concentration of filler on the
viscosity of a suspension. All of these equations predict the general
type of behavior shown in Figure 11-1. At low concentrations the
viscosity increases gradually with increasing filler concentration. As
the concentration increases further the curve becomes increasingly
steep, and at some limiting value of the concentration the viscosity
rises without bound.

103r-----------------~----------------------_,

~ 10 2

Ui

()
(fJ

::;:
W

>

i=

....l

w 10

ex:

1.0~~

__

0.1

____

0.2

____

0.3

____

____

0.4

____

0.5

____..J
0.7

VOLUME FRACTION

Figure 11-1. Sketch showing the relative viscosity as a function of volume fraction solids for a
typical suspension of particles in a Newtonian liquid.

RHEOLOGY OF MULTI PHASE SYSTEMS

393

This limiting concentration can be identified with the "maximum


packing fraction," m' of the filler. The concept of a maximum
packing fraction is simple in ideal situations. For example, if a
volume of space is packed with spheres arranged on a simple cubic
lattice, the fraction of the total volume occupied by the spheres is
(477'/3) /8 = 0.524. In real situations there is a distribution of
particle sizes, the particles are irregular in shape rather than
spherical, and the packing is random. It becomes more difficult,
therefore, to calculate the maximum packing volume precisely, but
estimates are available for a variety of cases [3, p. 136].
As an example of the many equations proposed we give that of
Maron and Pierce [8], because it has a simple form and has been
found to describe the effect of concentration for a variety of fillers.
TJ
TJ=-=
r
TJf

[1 - (/A)]2

(11-2)

The volume fraction of filler is , and TJ r is the relative viscosity of


the suspension. Kataoka et al. [9] found that setting A equal to 0.68
gave a good fit of data for suspensions of glass spheres in molten
polymers. This is close to the value of 0.637 for m for random
close packing of spheres [3, p. 136]. Kataoka and co-workers [10-12]
have also found the Maron-Pierce equation to apply to melts filled
with particles other than spheres. For rough irregular particles of
calcium carbonate, A was found to be 0.44. If we identify A with
m' this result makes sense, because it is not possible to pack
irregular particles as closely as spheres and because the effective
size of the filler particles is increased by a layer of adsorbed
polymer.
The viscosity of fiber-reinforced melts has also been found to be
represented by the Maron-Pierce equation [11]. In this case A (or
m) is found to decrease with increasing aspect ratio (length/diameter) of the fiber. It is interesting to note that even in this case the
experimental values of A are close to theoretical values of m for
randomly packed rods. A comparison is shown in Table 11-1. The
product of m and the aspect ratio is also shown; it is comparable
to the theoretical value for maximum isotropic packing of rod-like
molecules.

394

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Table 11-1. Parameter Values for Fiber-Reinforced Melts

LjD

A [11]

30
16
8

0.16
0.33
0.44

<Pm [3, p. 135]


0.173
0.303
0.476

5.2
4.8
3.8

A comment about the significance of the quantity <Pm for fiber


suspensions is in order. Recently a number of fiber-filled resins
have been introduced commercially that have fiber lengths as long
as one-half inch and with fiber loadings as high as 60 weight %
(> 40 volume %) [13]. The product of <Pm and LID far exceeds the
<Pm LID product predicted theoretically. This apparent contradiction is resolved by noting that the theoretical value of <Pm LID is
that for random packing. In these high-aspect-ratio, highly-Ioadedsystems the fibers are substantially aligned parallel to each other.
The processability of these systems is due, then, to the reduction of
viscosity caused by this alignment. This effect is superficially analogous to the reduction of viscosity caused by the liquid crystalline
ordering of rod-like molecules.
The Maron-Pierce equation is also consistent with the useful
observation that the viscosity of a concentrated suspension can be
reduced by using a filler with a distribution of sizes rather than one
of uniform particle size. Alternatively, such a filler can be used to
achieve a volume loading that exceeds the maximum value allowed
for the packing of uniform spheres. The concept is that the smaller
particles can fit into the interstices left by the packing of the larger
ones. Examples of the application of this idea can be found in the
literature [2]. As a specific example, the viscosity of a 60 volume %
suspension of spheres is reduced about 100-fold if 25% of the
spheres have a diameter 0.138 times that of the larger spheres.
Alternatively, such a mixture of sizes has the same viscosity at 74%
loading as does the uniform size suspension at 60% volume fraction. There is a practical limitation to the application of this
concept; if the smallest size fraction is too small it is subject to
agglomeration by colloidal forces and becomes difficult to disperse.
Some complicating factors concerning the viscosity of suspensions
need to be mentioned. First, for a suspension in a Newtonian fluid
the definition of the relative viscosity 1J r in Equation 11-2 is

RHEOLOGY OF MULTIPHASE SYSTEMS

395

105~--------------------------------~

10'

<0

e::>-

I-

U5
o
u
(j)

:> 10 3

102~

__

4.10 2 10 3

________

________

10'

________

10 5

10 6

SHEAR STRESS (Pa)

Figure 11-2. Viscosity versus shear stress for two suspensions of carbon black in LDPE at
150a C and for the unfilled melt. Top curve,S volume %; middle curve, 1 volume %. Adapted
from Ref. 14. Copyright 1985 by The Society of Plastics Engineers. Reprinted by
permission.

unambiguous. However, because polymer melts are generally nonNewtonian, the viscosity of the suspending fluid 171 depends on the
shear rate. The Maron-Pierce equation appears to be suitable if the
viscosity of the medium 171 and of the suspension are both measured at the same shear stress. This means that if the viscosities of
the filled and unfilled materials are plotted versus shear stress, as
shown in Figure 11-2 [14], the curves can be superposed by means
of a vertical shift. This, in turn, implies that if the viscosities are
plotted versus the shear rate, the flow curves will tend to converge
at high shear rates, as illustrated in Figure 11-3 [15]. This is a
fortunate circumstance, because processing operations are often
limited by the high shear rate viscosity, which is less strongly
affected by fillers than the low shear rate viscosity. A second
complicating factor is that high filler loading can cause an additional increase of the viscosity at low shear rates, as shown in
Figure 11-4 [15]. Such an upturn is often considered to be an
indication of a "yield stress" (J'o, a stress below which the material

396

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

10 6

r----------------------------------------------,

10 5
(j)
ell

>-

I-

Ui

10'

()
(/)

:>

10 3

______ ____
____
10 -3
10 2
10'

102~

~~

______

~~

10

____

______
10 2
10 3

SHEAR RATE (5-')

Figure 11-3. Viscosity versus shear rate for several loadings of TiO z in a HDPE. From
bottom to top, volume % = 0, 13 and 22. Note that the curves tend to converge at high
values of y. There is no evidence of a yield stress at these loadings. Adapted from Ref. 15.
Copyright 1976 by John Wiley & Sons, Inc. Reprinted by permission.

106~---------------------------------------------,

(j)
ell

>-

!:: 10'

(/)

()

(/)

:>

10 3

102~

10

____~~~__~______~~____~______~______~
10 '
1
10
10 3

SHEAR RATE (5-')

Figure 11-4. Viscosity versus shear rate for several loadings of Ti0 2 in a HDPE. The sharp
increase in viscosity at low shear rates suggests the presence of a yield stress. From bottom to
top, volume % = 0,4, 13,22 and 36. Adapted from Ref. 15. Copyright 1976 by John Wiley
& Sons, Inc. Reprinted by permission.

RHEOLOGY OF MULTIPHASE SYSTEMS

397

en
en
w

a:
~
a:
c(
w

J:

en
o~

__________________________________

SHEAR RATE

Figure 11-5. Typical shear stress versus shear rate behavior for a material having a yield
stress.

behaves as a solid and does not flow. Figure 11-5 shows a typical
curve of shear stress versus shear rate for steady simple shear of a
material having a yield stress. The yield stress has been found [9, 10]
to increase exponentially with volume fraction, cf>, of the filler.
Particles with diameters of 1 micron or less tend to be subject to
strong attractive forces due to the high surface area of, and small
distance between, particles. These forces promote aggregation and
the formation of the type of structure that can support a yield
stress. Tanaka and White [16] formulated a theory for concentrated
suspensions of spheres that predicts a yield strength, with U o
increasing with cf> and with interparticle force, and decreasing with
increasing particle size. Studies of latices have confirmed that a
yield stress is associated with small particles and large attractive
forces [17]. One also observes a yield stress in melts filled with small
particles of talc [18], titanium dioxide [15], calcium carbonate [10, 19],
and carbon black [19,20,21]. Coupling agents, on the other hand,
appear to reduce the yield stress [15, 18, 19, 22].
Many viscometers operate at fixed speeds, and it is not possible
to determine the yield stress by direct measurement, as any nonzero
speed will be associated with a shear stress greater than the yield
stress. If a yield stress is suspected, it is often helpful to plot
viscosity versus shear stress, as shown in Figure 11-6 [14]. In such a
plot the viscosity will rise without limit as the yield stress is
approached from above.

398

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


10 '

Ul
<ll

(l.

r-------------------------------,

10"

>-

t:

(f)

g
:>

10'

v LOPE
10 3

LOPE + 20 voir,

CaC0 3

eLOPE + 19 voir; Ti0 2


LOPE + 30 volc, Ti0 2
______ ______ ____

102~~

4.10 2 10 3

10 4

10 5

______

10 6

10 7

SHEAR STRESS (Pa)

Figure 11-6. Viscosity versus shear stress for several loadings of Ti0 2 or CaC0 3 in a LDPE
and for the base resin, all at 150C. The loadings are: 20% CaC0 3 , 19% Ti0 2 and 30%
Ti0 2 . The suspensions with the highest filler content appear to have yield stresses. Adapted
from Ref. 14. Copyright 1985 by The Society of :'lastics Engineers. Reprinted by
permission.

A possible explanation for the existence of a yield stress is that


the suspended particles form a strongly interacting network, which
must be disrupted before flow can occur. Formation of such a
network is very sensitive to chemical interactions. For example, it
has been shown [3, p. 147] that the presence of a small amount of
water binds filler particles suspended in a non-polar medium into
agglomerates. Conversely, dispersing agents that can couple to and
coat filler particles can prevent the formation of such aggregates.
The simplest model for the shear stress of a fluid with a yield
stress during steady shear flow is that of the "Bingham plastic":

(1l-3a)

RHEOLOGY OF MULTI PHASE SYSTEMS

399

or, dividing by y
(1l-3b)

For suspensions, the "Casson" equation is often found to give a


better fit of experimental data:

(11-4)
Aggregation of particles leading to pronounced changes in rheology can be induced by the shearing of suspensions, especially if the
suspending medium is non-Newtonian. A beautiful demonstration
of how a dispersion of spheres can be made to form oriented fibrils
by shearing is given by Michele et al. [23]. A practical application of
what is very likely the same effect is the increased elasticity of a
polymer melt resulting from dispersion of finely divided polytetrafluoroethylene (PTFE) particles that form submicroscopic fibrils
[24,25].
It should be noted that the measurement of the properties of
fiber filled melts poses special problems. Fibers are readily oriented
by flow fields, particularly in extensional or compressional flows.
This has in fact been utilized in an elegant fashion to design
extrusion dies to orient the fibers in any desired direction [26]. The
problem is that the rheology depends on the orientation of the
fibers. When loading a cone-and-plate rheometer, for example,
the test sample will be subjected to squeezing, which tends to orient
the fibers in the plane of the rheometer gap. In capillary or slit
rheometry there is a strong extensional flow as the material moves
from a large diameter reservoir into the capillary or slit, and this
promotes the alignment of the fibers in the flow direction. The
shear flow in the capillary, on the other hand, tends to rotate and
tumble the fibers, disrupting the orientation induced at the entrance [27]. This variation of orientation along the capillary may be
responsible in part for an apparent variation of viscosity with
capillary length. Of course these effects are not merely problems in
rheometry; they also affect the properties of parts made from fiber
filled polymers.

400

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The extensional viscosity of polymer melts is not strongly affected


by spherical fillers. Fibers, on the other hand, enhance the extensional viscosity enormously, even at low concentrations, in Newtonian fluids. The ratio of extensional to shear viscosity (Trouton
ratio) of fiber suspensions far exceeds the value of 3 expected for
Newtonian fluids. This finding is in good agreement with theory [2].
The elongation to break in melt extension experiments has been
found to decrease with increasing fiber content [28].
11.2.2 Elasticity

Our understanding of the elasticity of melts filled with rigid particles is considerably poorer than that of the viscosity. For unfilled
polymers the common measures of melt elasticity are the first
normal stress difference, the extrudate swell, and the capillary flow
entrance pressure drop. Although the quantitative relationships
among these measures are not known precisely, their trends with
changes of structure or conditions are generally in the same direction. This is not the case, however, with rigid particle filled melts.
The first normal stress difference is reduced by the presence of
small particles [15,19,22,29]. The curve of Nl versus a for a
polystyrene filled with large glass beads was found to be the same as
for the unfilled melt [30]. For fiber suspensions, Nl is found to be
increased [28]. However, there is disagreement as to whether this
increase is greater than or less than the increase of viscosity at
constant shear stress. The higher the stiffness and length of the
fiber, the greater is the increase in N l Suspensions of fibers even in
Newtonian fluids exhibit a nonzero value of N l .
There is no clear pattern of the effect of fibers on the Bagley end
correction for capillary flow [31]; both increases and decreases with
respect to the unfilled melt have been observed. The addition of
more than 5% filler to low density polyethylene suppressed vortices
at the entrance to a die having a 1800 (flat) entrance angle [14]. The
same loading of filler also resulted in a viscosity versus shear stress
curve having a shape suggesting the presence of a yield stress.
Coupling agents appear to reduce the yield stress [15,18,19,22].
The addition of both reinforcing [31,32] and nonreinforcing
[15,32] fillers significantly reduces extrudate swell. For example, the
addition of 30% titanium dioxide to high density polyethylene

RHEOLOGY OF MULTI PHASE SYSTEMS

401

reduces swell by 65%. Furthermore, it has been known for many


years that the addition of carbon black to raw rubber, in the
manufacture of tires, substantially reduces the elasticity of the
elastomer and makes it much easier to process. The mechanism by
which fillers inhibit swell is not understood at this time. There is
one notable exception to this observation. When the filler is a long
fiber and the die length is very short, extremely high extrudate
swell, and even foaming, is observed [33-35]. This effect has been
attributed to recovery of the elastic deformation of the fibers
generated at the entrance to the die. The high extrudate swell of
the PTFE-fibril filled melts mentioned above is probably due to the
same mechanism.
11.3 DEFORMABLE MULTIPHASE SYSTEMS (BLENDS, BLOCK POLYMERS)

Immiscible blends of polymers, in which two (or more) phases


coexist, are of increasing commercial importance. Their rheological
behavior differs from that of homogeneous melts, as do those of
melts containing rigid fillers, because the flow field is affected by
the presence of a second phase. The simplest situation is a dilute
suspension of a Newtonian fluid in a Newtonian matrix. In the limit
of very low shear rates, where the suspended fluid drops retain
their spherical shape, G. 1. Taylor derived an equation for the
viscosity of the suspension that is analogous to the Einstein relation
(11-1) for rigid particles:
7J = 7Jf [ 1

1 + 2.5 P
+ ( 1 +P

tjJ

(11-5)

where P is the ratio of the viscosity of the disperse pbase 7Jd to that
of the matrix phase 7Jf. When P becomes very large the equation
for 7J reduces to the Einstein relation. However, even if p goes to
zero, as in the case of a foam, Equation 11-5 predicts that the
viscosity is higher than that of the suspending fluid.
Suspensions of one liquid in another are more complicated than
dispersions of rigid particles, however, because of the deformability of the disperse phase. Flow causes the suspended "particles"
to change their shape and thus their effect on the flow. This be-

402

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

havior, along with the effect of interfacial tension, introduces nonNewtonian and elastic effects such as the shear rate dependence of
viscosity and nonzero values of Nt, even for suspensions of Newtonian fluids [36].
Furthermore, neither the size of the suspended phase regions nor
their "morphology" are constant but depend on the flow conditions.
By morphology we mean the geometric type of the suspension; the
suspended phase may be in the form of isolated drops or fibrils or
have a stratified layer structure, or the suspension may consist of
interwoven networks in which both phases are continuous. The size
and morphology are determined by the balance of the viscous
forces, which depend on the rheology of the components, and by
the interfacial tension. The processes involved are the deformation
of the disperse phase, which may lead to break-up, and the coalescence of colliding droplets or fibers.
The importance of interfacial tension, even in viscous polymer
melts, is easily seen by considering the stresses involved. Across a
curved boundary between two immiscible fluids there is a pressure
difference on the order of s / a, where s is the interfacial tension
and a is the radius of curvature. For a drop of fluid with a tension
of 10 dynes/cm and a radius, a, of 1 micron (10- 4 cm) the pressure
difference is about 10 5 dynes/cm 2 This corresponds to the shear
stress in a melt with a viscosity of 1000 Poise at a shear rate of
100 S-I.
Just as in rigid particle suspensions, the interaction of disperse
phase domains in concentrated systems makes it impossible to
predict quantitatively their behavior by extrapolation of the dilute
suspension behavior. Nevertheless, studies on dilute systems give
insight into the mechanisms involved and the effects of the relevant
factors. The results of such studies are summarized in the following
section and compared with experimental measurements of polymer
blends.
The rheology of immiscible blends is quite complicated, as can be
concluded from the above discussion, and there seem to be few
generalizations that can be made. A very brief summary is given in
the subsequent section. Phase-separated block copolymers represent a type of immiscible blend, with the added restriction that the
two components are chemically bonded. Their behavior is also
summarized in that section.

RHEOLOGY OF MULTIPHASE SYSTEMS

403

11.3.1 Deformation of Disperse Phases and Relation to Morphology

At rest the drops of fluid suspended in a matrix are spherical,


because this shape minimizes the interfacial energy. In flow the
drop shape is deformed, initially to an ellipsoid whose major axis is
oriented at an angle to the flow field. The shape of the ellipsoid can
be described in terms of a parameter D, which is related to the
lengths of the major and minor axes of the ellipsoid, Land B:

D=

L-B
L+B

(11-6)

Taylor showed that D is governed by the ratio of the viscous stress


71f1 exerted on the drop to an interfacial stress, s/a, where s is the
interfacial tension, and a is the radius of curvature of the drop:
D

171f a

== We

(11-7)

The dimensionless group on the right-hand side of this equation is


known as the Weber Number (We) and is an important parameter
for characterizing disperse phase size and morphology. Equation
11-7 accounts for the fact that a shear field tends to elongate the
drop, whereas the elongation is resisted by the interfacial tension,
which tends to minimize the surface area increase caused by the
elongation. Also, it shows that a large drop is more easily deformed
than a small one.
Taylor also showed that when D (or We) exceeds some critical
value, the deformed drop breaks up into smaller ones. The critical
value of We depends on the viscosity ratio 71d/71f and on the type
of flow field. The nature of the breakup process also depends on
these variables [37,38]. The deformed drops can neck and break up
into roughly equal sized drops; these may have pointed ends from
which tiny droplets are expelled, or they may be drawn out into
long fibrils that eventually break up.
Experimental results for the dependence of We on 71d/71f at
breakup are shown in Figure 11-7. An extensional flow field is very
effective in causing rupture at all values of 71d/71f' In shear flow,
however, breakup occurs most readily when 71d/71f is close to unity,

404

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


100~----------------------------~--------------~

[
0::

III

:2

'" "

:l
Z
0::

w
w

III

7------- ,,""

---............

/
~

NEWTONIAN SHEAR

----------------------~------I
NEWTONIAN ELONGATIONAL

0.1~~~~. .~~~~~~~~~~~. .~__~~~~~

10

10 '

10

10 2

VISCOSITY RATIO (TJd/TJ,)

Figure 11-7. Weber number at breakup as a function of TJd/TJf for a suspension of one
Newtonian fluid in another, for shear flow and for extensional flow. Also shown is a curve for
blends of molten polymers in an extruder. Adapted from Ref. 43. Copyright 1987 by The
Society of Plastics Engineers. Reprinted by permission.

i.e., when the matrix fluid and the disperse phase have nearly the
same viscosity. When 1Jd/1J! exceeds a value of between 3 and 4,
the matrix does not exert sufficient stress to cause rupture. When
1Jd/1J! is very small, less than about 0.005, the suspended drops are
readily drawn out into long fibrils but do not break up. For a given
system the conditions for breakup in a shear flow can be predicted
theoretically from the extensional flow breakup criteria [39].
The breakup of suspended drops is also affected by other factors.
In an unsteady shear flow, in which the shear rate varies with time,
both the mechanism of drop rupture and the critical conditions
depend on the rate of increase of the shear rate [38]. Such a rate
dependence is perhaps to be expected since the interfacial tension
acts as an elastic restoring force. The presence of emulsifying
agents has also been found to affect breakup.

RHEOLOGY OF MULTIPHASE SYSTEMS

405

All of the studies referred to so far were done with inelastic


fluids. Van Dene [40] suggested that the effect of viscoelasticity is to
stabilize drop shape if the disperse phase is more elastic, i.e., has a
higher elastic compliance, than the suspending fluid. Van Dene
added a factor proportional to the difference of the normal stresses
of the two fluids to the interfacial tension. This hypothesis has
recently been confirmed at least approximately [41].
Elasticity also affects the shape and the rate of breakup of the
long fibrils or threads that can be drawn out from drops [42]. This
may be due in part to the high extensional viscosity of viscoelastic
fluids, because the breakup process requires some extensional flow.
An even more significant factor influencing the breakup of threads
is the existence of a yield stress, which tends to stabilize the threads
against breakup.
The above studies on isolated or dilute drops indicate that the
Weber Number and the ratio 7Jd/7Jf characterize the size of the
disperse phase when immiscible fluids are deformed. These parameters have also been found to correlate data for concentrated
blends of polymers [28]. However, the actual disperse phase size for
concentrated suspensions is affected by concentration, being larger
at high concentrations of the disperse phase. This is because the
small drops formed by the breakup process can coalesce to form
larger drops [44]. The coalescence process depends strongly on the
nature of the interface between the drops.
The addition of a third phase consisting of a block polymer of the
blend components has been shown to be effective in reducing
domain size [45]. This effect has been attributed to a reduction of
the interfacial tension [43] or to an effect on the interface mobility
[44].
It is important to remember that the blend morphology that
results from commercial processing is difficult to quantify, because
of the complex flow field involved and because of the transient
nature of processing. As shown above, shear and elongational fields
differ in their ability to cause breakup or drawing out of fibrils, and
polymer processing operations generally involve a mixture of these
flows. Also, in an unsteady flow the breakup process depends on
the rate of change of the deformation rate. And not least, the
breakup and coalescence processes do not occur instantaneously.

406

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

For viscous polymer melts the times required for these processes
may exceed the process residence time, so that the resulting morphology is not an equilibrium state [42]. This time dependence
appears to account for the formation of a co-continuous morphology, because drawn-out threads have not had time to break up
during processing. The remarkable ability of phase-separated block
copolymers to form such a morphology [42, 46] is consistent with
this view, because these polymers have yield stresses that stabilize
their fibrils.
One last point that should be made here is that the stresses
resulting from flow can affect the conditions for miscibility of two
polymers. This was shown some time ago for polymer solutions.
More recently it has been found that the phase separation temperature of polymer blends also changes with flow, both in shear [47]
and in extension [48].
11.3.2 Rheology of Immiscible Polymer Blends

In light of the interdependence of deformation history and morphology, and therefore also of rheological behavior [49], it is not
surprising that few generalizations can be made about the rheology
of blends of immiscible polymers. Because of the diversity of results
the numerous studies cannot be briefly reviewed, and the reader is
referred to the book by Han [4] for a detailed discussion.
The melt viscosity of series of blends of different concentrations
of a given pair of polymers may be intermediate between that of the
components. In many cases, however, it is found that the viscosity
goes through a maximum or a minimum as a function of concentration. Compared at constant stresses, the viscosity may go through a
maximum, higher than either component, at low stress, and then
through a minimum, lower than that of either component, at high
stress. The normal stress difference appears to have the opposite
behavior, i.e., it goes through a maximum when the viscosity shows
a minimum, and vice versa.
The existence of a viscosity minimum is difficult to understand. In
some cases of stratified (two layer) flow a minimum can occur under
special circumstances, when the flow curves (viscosity versus shear
stress) of the two components cross [4]. It is not clear that this effect
can account for the observed minima in disperse blends. Another

RHEOLOGY OF MULTI PHASE SYSTEMS

407

possible explanation is that lubrication of the capillary wall can


cause apparent slippage and increase the flow rate [50].
The rheology of polymer blends is further complicated if a
chemical reaction can occur between the blend components, as for
example with polyesters and polyamides [51]. Such a reaction can
produce block or graft polymers, whose effect on the reduction of
disperse phase size has been discussed above.
At this time the only definite recommendation that can be made
is that the rheology of blends of interest should be measured under
conditions (temperature, flow field type, shear rate or stress range,
residence time) approximating as closely as possible the conditions
anticipated during processing. And of course the preparation of the
blends themselves should be done under realistic compounding
conditions.
11 .3.3 Phase-Separated Block and Graft Copolymers

Block and graft copolymers in which the homopolymers of the


components are immiscible will form phase-separated systems if the
block lengths are sufficiently long. These are therefore analogous to
the immiscible blends discussed above with the additional restriction that the component phases are chemically bonded. This restriction imposes a constraint on the phase morphology.
A detailed study of one such polymer, styrene-butadiene-styrene
block copolymer, has been conducted by a working party of IUPAC
[52]. The observations of this study are generally similar to the few
previously published results and are summarized below.
The viscosity of the block copolymer is higher than that of a
random copolymer of comparable molecular weight. At low shear
rates there is an increase of viscosity, apparently due to a yield
stress. Particularly at low shear rates the rheology is sensitive to the
deformation history, and results are poorly reproducible and depend on details of sample preparation. The flow does not reach a
steady state even after long shearing times, and there is a large
effect of the "rest time" between successive measurements at
different shear rates.
The Cox-Merz rule does not hold for the block copolymer.
Normal stress measurements are not reproducible, and the data
show multiple maxima as a function of time of shearing. In one case

408

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

a negative first normal stress difference has been observed. Extrudate swell is reduced compared to that in homogeneous polymers.
The extensional viscosity increases rapidly with decreasing extension rate below 10 s -1, more rapidly than the increase of the shear
viscosity.
The temperature dependence of viscosity is dominated by that of
the styrene block component, which is closer to its Tg and therefore
more temperature dependent than the butadiene block. These
results are consistent with a picture in which a network structure
exists at rest or at low shear stresses. High stresses disrupt this
structure and decrease the viscosity. The flow curve therefore
exhibits shear thinning, and its shape resembles that of a material
with a yield stress. Transient effects result from the slow restoration
of the network structure.
REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.

D. c.-H. Cheng, Powder Technology 37:255 (1984).


A. B. Metzner, 1. Rheol. 29:739 (1985).
L. E. Nielsen, Polymer Rheology, Marcel Dekker, N.Y., 1977.
C. D. Han, Multiphase Flow in Polymer Processing, Academic Press, N.Y., 1981.
J. L. White and K. Min, in Polymer Blends and Mixtures, ed. by D. J. Walsh,
J. S. Higgins and A. Maconnachie, Martinus Nijhoff Publishers, Boston, 1985.
M. R. Kamal and A. Mutel, 1. Polym. Eng. 5:293 (1985).
S. A. Khan and R. K. Prud'homme, Reviews in Chem. Eng. 4:205 (1987).
S. H. Maron and P. E. Pierce, 1. Coli. Sci. 11:80 (1956).
T. Kataoka, T. Kitano, M. Sasahara and K. Nishijima, Rheol. Acta 17: 149
(1978).
T. Kataoka, T. Kitano, Y. Oyanagi and M. Sasahara, Rheol. Acta 18:635
(1979).
T. Kitano, T. Kataoka and T. Shirota, Rheol. Acta 20:207 (1981).
T. Kitano, T. Kataoka and Y. Nagatsuka, Rheol. Acta 23:20 (1984).
B. Miller, Plastics World, December, 1986, p. 28.
c.-Y. Ma, J. L. White, F. C. Weissert and K. Min, SPE Tech. Papers 31:131
(1985).
N. Minagawa and J. L. White, 1. Appl. Polym. Sci. 20:501 (1976).
H. Tanaka and J. L. White, 1. Non.-Newt. Fl. Mech. 7:333 (1980).
1. M. Krieger, Advances in Colloid Science 3:111 (1970).
F. M. Chapman and T. S. Lee, SPE loum. 20:37 (1970).
H. Tanaka and J. L. White, Polym. Eng. Sci. 20:949 (1980).
G. V. Vinogradov et aI., Int. 1. Polym. Mat. 2:1 (1972).
V. M. Lobe and J. L. White, Polym. Eng. Sci. 19:617 (1979).

RHEOLOGY OF MULTIPHASE SYSTEMS

22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.

409

C. D. Han, C. Sandford and H. J. Yoo, Polyrn. Eng. Sci. 18:849 (1978).


J. Michele, R. Piitzold and R. Donis, Rheol. Acta 16:317 (1977).
W. F. Busse, J. Polyrn. Sci., A-2 5:1249 (1967).
W. F. Busse and G. H. Bowers III, U. S. Patent 3,005,795; Oct. 24, 1961.
L. A. Goettler, R. I. Leib and A. J. Lambright, Rubber Chern. Technol.52:838
(1979).
R. J. Crowson, M. J. Folkes and P. F. Bright, Polyrn. Eng. Sci. 20:925 (1980).
Y. Chen, J. L. White and Y. Oyanagi, J. Rheol. 22:507 (1978).
J. L. White, L. Czarnecki and H. Tanaka, Rubber Chern. Technol. 53:823
(1980).
K. Oda, J. L. White and E. S. Clark, Polyrn. Eng. Sci. 18:25 (1978).
Y. Chan, J. L. White and Y. Oyanagi, Polyrn. Eng. Sci. 18:268 (1978).
T. Nishimura and T. Kataoka, Rheol. Acta 23:401 (1984).
R. J. Crowson and M. J. Folkes, Polyrn. Eng. Sci. 20:934 (1980).
M. J. Folkes, Short Fiber Reinforced Thermoplastics, John Wiley & Sons, N.Y.,
1982.
S. Turner and F. N. Cogswell, Proc. VIIth Internat. Congr. Rheo!., p. 172,
Gothenburg, 1976.
C. D. Han and R. G. King, J. Rheol. 24:213 (1980).
F. D. Rumscheidt and S. G. Mason, 1. Coli. Sci. 16:238 (1961).
S. Torza, R. G. Cox and S. G. Mason, J. Coli. Interface Sci. 38:395 (1972).
W. L. Olbricht, J. M. Rallison and L. G. Leal, J. Non-Newt. Fl. Mech. 10:291
(1982).
H. J. Van Oene, J. Coli. Interface Sci. 40:448 (1972).
J. J. Elmendorp and R. J. Maalcke, Polyrn. Eng. Sci. 25:1041 (1985).
J. J. Elmendorp, Polyrn. Eng. Sci. 26:418 (1986).
S. Wu, Polyrn. Eng. Sci. 27:335 (1987).
J. J. Elmendorp and A. K. Van Der Vegt, Polyrn. Eng. Sci. 26:1332 (1986).
R. Fayt, R. Jerome and Ph. Teyssie, Polyrn. Eng. Sci. 27:328 (1987).
W. P. Gergen, S. Davison and R. G. Lutz, Rubber Chern. Tech. 58:857 (1985).
K. A. Mazich and S. H. Carr, J. Appl. Phys. 54:5511 (1983).
J. D. Katsaros, M. F. Malone and H. H. Winter, Polyrn. Bulletin 16:83 (1986).
A. P. Plochocki, Polyrn. Eng. Sci. 23:618 (1983).
C. K. Shih, Polyrn. Eng. Sci. 16:11 (1976).
L. A. Utracki, A. M. Catani, G. L. Bata, M. R. Kamal and V. Tan, J. Appl.
Polyrn. Sci. 27:1913 (1982).
A. Ghijsels and J. Raadsen, Pure Appl. Chern. 52:1359 (1980).

Chapter 12
Chemorheology of Reacting
Systems

12.1 INTRODUCTION

The term chemorheology was first introduced by Tobolsky et al.


[1,2] to describe their research on the "chemical stress relaxation"
of cross-linked rubbers. They found that the stress in a stretched
specimen decayed to zero over a long period of time, a behavior
incompatible with the concept of a cross-linked structure. They
concluded that cross-links had been gradually lost in the strained
sample so that the apparent relaxation was actually due to a
chemical change and was not a viscoelastic effect. More recently,
the term has come to be used to describe the study of rheological
changes occurring during the course of any chemical reaction.
Rheological properties are commonly used to monitor degradation processes in polymers [3,4]. Indeed, a common procedure for
evaluating useful sample life and the effectiveness of stabilizers is to
monitor the storage and/or loss moduli over an extended period of
time at an elevated temperature. However, the major focus of
attention at the present time is the effect of cross-linking reactions
on rheological properties [5], and that is the subject of this chapter.
Cross-linking reactions are obviously of central importance in the
rubber industry, but cured elastomers are outside the scope of our
treatment.
The rheology of cross-linking reactions is of interest in several
contexts. First, rheological measurements are often used to monitor
the extent of a curing reaction in the laboratory. Other analytical
techniques used for this purpose include differential scanning
calorimetry (DSC), high pressure liquid chromatography (HPLC),
410

CHEMORHEOLOGY OF REACTING SYSTEMS

411

and infrared spectroscopy (IR), but rheological properties are more


sensitive to the transition from the liquid to the gel state. Rheological properties are also of interest because of their importance in
manufacturing processes involving curing reactions. In particular, it
is crucial that gelation not occur until all necessary flow processes
have been completed. Thus, rheological properties are used in
material evaluation, process design, and process simulation.
Cross-linked polymers are widely used as structural components
in the aerospace industry [6,7] and as adhesives for aerospace and
other applications [8]. They are also used in the electronics industry, in the lamination and packaging of multilayer circuit boards
[9, 10], and for solder mask coatings [11]. In addition to these
specialized applications, a wide range of articles are now made of
thermosetting materials using conventional molding processes [12]
as well as by reaction injection molding (RIM) [13-18]. Another
important application of cured polymers is in low-solvent coatings
[19-22], where the viscosity plays a central role in the leveling
process.
12.2 NATURE OF THE CURING REACTION

Once all reactants are mixed and the temperature is sufficient for
the cross-linking reaction to proceed, polymerization will occur by
one of several mechanisms depending on the chemical nature of the
system. At first, branches will increase in number, and there will be
a corresponding increase in molecular weight and viscosity. Branches
connecting two molecules become cross-links, and at a level of
about one cross-link per molecule, a three-dimensional network or
gel is established that prevents further flow [23]. The time at which
this occurs is termed the gel point. The gel is a rubbery material,
but usually still contains some soluble polymer. Further curing will
lead to vitrification, i.e., the conversion to a rigid glassy material.
While this is the normal progression of events in the processing
of a thermosetting polymer, a rich field of possible phase changes
emerges if we consider a very broad range of curing temperatures.
The various possibilities can be seen clearly by reference to a
time-temperature-transformation (TTT) isothermal cure diagram of
the type proposed by Enns and Gillham [24]. Such a diagram for a
cross-linking system is shown in Figure 12-1. For a given curing

412

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

,
",,,

Tgx

LU

a:
=>

. "

Sol/gel
rubber

,,
,

EIClstOJl1pr

,
~~~~~~~~~~

a:

LU
Q.

~
LU
~

Cl
Z

a:
=>

LOG TIME

Figure 121. Timetemperature-transformation isothermal cure diagram of the type proposed


by Enns and Gillham. Adapted from Ref. 24. Copyright 1983 by John Wiley & Sons, Inc.
Reprinted by permission.

temperature, moving horizontally to the right shows the various


transitions that will occur. The curves shown in the liquid region
are lines of constant viscosity. The curing temperature for thermosetting systems is between geJTg and Tgx.' while in rubber vulcanization, it is above Tgx..
A commercial thermoset molding resin is often stored and shipped
as an ungelled glass. Upon heating to the curing temperature, this
is converted first to a liquid, then to a gel and finally to a gelled
glass.

CHEMORHEOLOGY OF REACTING SYSTEMS

413

If phase separation is also occurring, as in rubber-modified mate-

rial, this will alter the picture somewhat, as the processes of phase
separation, gelation and vitrification affect each other [15,24].
12.3 EXPERIMENTAL METHODS FOR MONITORING CURING REACTIONS

Rheological measurements supplement the information available


from other techniques such as gel permeation chromatography,
differential scanning calorimetry, infrared spectroscopy and dielectric analysis [6, 25]. Rheometry provides the most direct measure of
the overall progress of the reaction and the approach to the gel
point. While dielectric analysis is a more convenient method of
tracking certain stages of curing reactions, this technique does not
yield data on flow properties that are essential for engineering
design and process simulation.
The main difficulties that arise in the use of a rheometer to
monitor a curing reaction are as follows.
1. Maintaining a constant, uniform temperature is much more
difficult than in a nonreactive system. The sample must be
heated to the curing temperature after it is placed in the
rheometer. This will take some time, during which the viscosity will fall and some reaction will occur. If the reaction is a
fast one, this will make it difficult to carry out an isothermal
measurement. If the reaction is exothermic, this will exacerbate the problem of maintaining a uniform temperature in the
sample.
2. In order to make a rheological measurement, it is essential
that the sample adhere to the walls of the rheometer. However, once curing has occurred it is often very difficult to
remove the sample. This suggests the use of disposable
rheometer fixtures.
3. During pregel curing, the viscosity may increase by several
factors of ten. Using a single shear rate and torque transducer,
it is generally not possible to track the entire increase up to
the gel point. Furthermore, once gelation occurs, flow is no
longer possible, and viscosity is no longer a useful property for
tracking the process. Thus, it is a challenge to devise a single
measurement technique that can track the entire curing pro-

414

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

cess from monomer to fully cured thermoset. As a result, as


many as three types of instrument have been used to track a
single curing reaction [26].
4. Slip and edge flow irregularities are more likely to occur, and
the gel can be destroyed if the strain is too large.
Empirical resin evaluation methods avoid some of these problems
but do not yield well-defined physical properties. For example,
spiral flow molds [27-30] can be used to compare several resins in
terms of the flow time available for gelation. ASTM 1 Method
D3123 [28] is a spiral flow test that is widely used, particularly in the
semiconductor encapsulation industry [29]. However, there is no
direct relationship between test results and the moldability problems that arise in manufacturing [30].
In the "monohole flow test" of Tonogai and Seto [31] high
frequency heating is used to preheat the sample, and a simple flow
geometry is used to improve the degree of control over the experimental variables. A related technique makes use of a "dynamic
extrusion rheometer," in which an orifice plate is forced down in a
reservoir containing the sample, which then flows up through the
orifice [32].
Because of their general popularity and availability, methods
have been proposed to use capillary rheometers to measure viscosity up to the gel point. The clean-up problem has been solved by
the use of disposable steel inserts [33] or by the use of an actuatordriven piston to clear the barrel and capillary [34].
The squeezing flow generated between two parallel plates by a
servohydraulic testing system has been used to evaluate epoxy resin
used in the manufacture of multilayer electrical circuit boards
[9, 10]. Using the analysis for a Newtonian fluid [35] the viscosity
can be calculated from the force during curing.
Because of their versatility, rotational rheometers have proven
the most useful for monitoring curing reactions. In a constant speed
mode of operation, the viscosity can be measured, while oscillatory
shear yields the components of the complex modulus. Constant
torque instruments have also been used to advantage. Concentric
lAmerican Society for Testing and Materials, Philadelphia, P A.

CHEMORHEOLOGY OF REACTING SYSTEMS

415

cylinder, cone-plate, and parallel disk fixtures have all been used,
although the latter two geometries are usually preferred, as they are
more easily cleaned. The eccentric rotating disk (ERD) geometry
has also been used to study curing [36].
General purpose rotational rheometers have been widely used to
monitor the viscosity and the complex modulus [8,37,38]. ASTM
Method D4473 [39] is based on the use of a cone-plate rheometer
operated in the oscillatory mode. There have also been attempts to
measure the first normal stress difference [12], but because curing
reactions are generally accompanied by a change in volume, it is not
possible to make a quantitative determination of this property. In
the case of foam [40], even the shear stress loses its normal
significance, and only "apparent" properties can be determined.
In order to accommodate the large change in viscosity that
occurs, a sequence of fixtures of decreasing radius [37] or a sequence of decreasing shear rates [41] can be used to track the
various stages of the reaction. Another approach to this problem is
to operate in the oscillatory shear mode and decrease the strain
amplitude as the curing proceeds, to keep the torque within the
range of the transducer.
Constant torque rheometers are useful in the monitoring of the
pre-gel stage of the reaction, as the torque can be set to a fixed
level and the speed recorded as a function of time to determine the
viscosity [15,42]. Disposable fixtures of both the cone plate [42] and
concentric cylinder [15] types have been used. The motion will stop
at some point after gelation starts to occur. Choy and Plazek [43]
used a sophisticated torsional creep apparatus to track the progress
of an epoxy cure by doing creep and recovery experiments. They
found that for viscosities above 10 4 Pa s it is not possible to
distinguish between permanent viscous deformations and recoverable contributions to strain.
In order to obtain fundamental property information, it is desirable to make measurements at a uniform, constant temperature. If
calorimetric measurements can be made at the same time, reaction
rate information and rheological properties can be obtained
simultaneously [44-46].
Isothermal operation becomes more and more difficult as the
reaction speed increases. An analysis of the heat transfer occurring
during the test becomes necessary to ensure that the assumption of

416

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

isothermality is valid [15,42]. For very rapid RIM reactions, for


which the gel time can be less than one minute, it becomes virtually
impossible to conduct isothermal experiments. Blake et al. [15] have
argued that in this case, it is the adiabatic process that is most
relevant to the actual RIM operation. Perry et al. [13] designed a
special on-line rheometer to study very fast curing reactions. The
parallel disk test fixtures are fed from a center hole in the stationary disk by a RIM machine. A surge cavity is located between the
RIM machine and the rheometer to hold excess material, as a
practical shot size provides more liquid than is necessary to fill the
rheometer. This cavity is also designed to prevent the first, atypical
portion of the shot from reaching the rheometer.
Another approach to the general problem of monitoring curing
reactions by use of mechanical measurements is the use of supported samples. In this technique, a spring or braid is coated with
liquid reactant and set into vibration. Torsional braid analysis
(TBA) [47] and dynamic spring analysis (DSA) [48] have been used
in this way. In the latter, a coated spring is subjected to a sinusoidal
force. 2 Of course, the support makes a major contribution to the
response of the system, and the resolution of the contribution due
to the sample is not straightforward [49]. Senich et al. [50] developed a model for the DSA technique that allows one to estimate
the storage and loss moduli for a resin at several frequencies.
However, this technique does not provide reliable data during the
advanced stages of the curing process due to the very small displacements that are generated as the sample vitrifies. The TBA
method is not subject to this limitation, but the interpretation of the
data is not as clear. Another technique involving a supported
sample is torsional impregnated cloth analysis (TICA). This technique makes use of an actual industrial material, subjecting it to
forced torsional deformation in a rotational rheometer equipped
with special fixtures [51-53].
Dynamic mechanical analysis (DMA) techniques, using unsupported samples, have proven useful for studies of the advanced
stages of a curing reaction [50,51,54] and of the effect of moisture
on epoxies [55].

2An instrument of the type used for dynamic mechanical analysis can be used to generate the
force. The Rheovibron has usually been used in the DSA method.

CHEMORHEOLOGY OF REACTING SYSTEMS

417

12.3.1 Dielectric Analysis

For a number of years, changes in the dielectric properties of a


material have been used to monitor curing reactions [6,56,57], and
it has been proposed that these changes can be related to chemical
and rheological changes. The use of dielectric analysis to track
cross-linking reactions has grown especially rapidly since the development by Senturia and co-workers of a micro-dielectrometer incorporating a capacitor, a differential amplifier and a temperature
sensor, all contained on a single small chip [58,59]3. This device can
be embedded in solid surfaces for the in situ measurement of
dielectric properties [59]. A dielectric sensor responds to the mobility of the ions present in some monomers, for example epoxy, due
to the synthesis reactions. This mobility is, in turn, a function of
molecular segment mobility, which changes substantially during
curing.
A question of importance for the interpretation of data is the
relationship between the dielectric loss factor and the viscosity
[6,60,61]. Gotro and Yandrasits [60] mounted a microdielectrometer in one disk of a parallel disk rheometer so that the dielectric
loss factor and the viscosity of an epoxy during curing could be
measured simultaneously. Both disks were made of aluminum and
were disposable. During the initial heating process, the loss factor
rises and the viscosity decreases. Then, as the cross-linking reaction
proceeds, the loss factor goes through a maximum and decreases,
while the viscosity goes through a minimum and increases. The loss
factor continues to decrease, approaching a constant value in the
gelled material, while the viscosity goes to infinity. Gotro and
Yandrasits [60] found that the maximum in the loss factor always
occurred somewhat later than the minimum in the viscosity, and
they discuss possible reasons for this.
A related technique involves the use of a frequency dependent
electromagnetic sensor (FDEMS). This technique has been described by Kranbuehl et al. [62] and used by Wetton et al. to make
simultaneous viscoelastic and dielectric measurements [63].4
3A laboratory instrument incorporating such a sensor is the DuPont 2970 Dielectric Analyzer.
A commercial version of the micro-dielectrometer is manufactured by Micromet Instruments
Inc. of Cambridge, MA
4This study made use of a commercial instrument, the PL-DETA, manufactured by Polymer
Laboratories, See Appendix E.

418

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

12.4 VISCOSITY OF PRE-GEL LIQUID

The viscosity of the monomer first decreases as it is brought to the


curing temperature. After going through a minimum it then increases as the reaction progresses. At the gel point flow is no longer
possible, and the viscosity thus becomes infinite. However, if a
rheometer is operated in the constant speed mode, a finite, fluctuating stress will be observed after the gel point, as the strength of
the initial gel is low, and it can be easily broken up into fragments
that are then tumbled if shearing is continued.
In general [42], the viscosity is a function of the temperature, the
shear rate, and the reactive group conversion, X.

YJ=YJ(T,X,y)

(12-1)

Newtonian behavior is usually observed in the pre-gel liquid unless


a solid phase is present. Therefore:
(12-2)
Restricting our attention to Newtonian behavior, or to the zero
shear viscosity of a non-Newtonian fluid, Equation 12-1 becomes:

YJ = YJ(T,X)

(12-3)

During an isothermal curing reaction, X increases with time, and


the observed variable will be the viscosity as a function of time, with
the temperature as a parameter that is constant during a given
experiment:

YJ = YJ(t, T)

(12-4)

The effect of temperature arises from the dependencies of both


viscosity and reaction rate on temperature. A simple, empirical
equation [41,64,65] that incorporates these dependencies is:

In[ YJ(t, T)]

In( YJoo) + flE1)/RT + tkoo exp(flEdRT) (12-5)

where YJoo' koo, flE1) and flEk are constants.

CHEMORHEOLOGY OF REACTING SYSTEMS

419

For a nonisothermal reaction the viscosity depends not simply on


the instantaneous temperature but on the temperature history.
Roller [10, 66] has generalized Equation 12-5 for this case to give:

In 77[t, T{t)]

In{77oo) + flE",/RT +

{koo
exp{flEk/RT) dt
o
(12-6)

This empirical model assumes that the reaction mechanism is


independent of the degree of reaction, but this may not be true, for
example when the reaction is diffusion controlled [68]. Keenan
[10,67,69] introduced a constant in the integral term to account for
an "entanglement effect," while Dusi et al. [69,70] introduced a
reaction order constant. Other attempts to model curing reactions
[68,71-73] have made use of the WLF equation rather than an
Arrhenius type term to account for the direct effect of temperature
on viscosity.
It is of interest to relate reaction rates and molecular weight to
viscosity [8, 10,42,44,68,74,75]. However, this procedure is complicated by the dependence of viscosity on temperature and by polydispersity. While reaction kinetics involves the number average
molecular weight, the viscosity is more sensitive to the weight
average molecular weight [74]. Using a novel "Reaction
Rheometer,"Biesenberger and Kumar [45] measured viscosity while
obtaining calorimetric data so that two independent measures of
the extent of reaction could be determined independently.
12.5 THE GEL POINT AND BEYOND

In theory, at the gel point the viscosity becomes infinite and the
material stops flowing and becomes an elastic gel, but one for which
the equilibrium modulus is zero. However, neither of these properties provides a useful experimental criterion for determining the gel
time for a specific resin. This is because when the cross-link density
becomes just sufficient for formation of a three-dimensional network, the structure formed is very weak. Thus, it can be easily
broken up by continued shearing. Furthermore, it is not practical to
detect the modulus of the gel just after it is formed.

420

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

A number of more or less empirical criteria have been proposed


to identify a gel point from rheological data [37,42]. For example,
the time at which 7]' (or 7]) = 10 3 Pa s is sometimes taken as the gel
point. Tung and Dynes suggested that G' = Gil at the gel point [76].
ASTM Standard Test Method D4473-85 [39] mentions this criterion
as well as two others. It suggests that the time when tan(o) reaches
its maximum value be taken as the "gel time," while the time at
which the storage modulus levels out to a constant value be called
the "cure time."
Winter and Chambon [77] pointed out that in order for the
criterion G' = Gil to be independent of frequency, it must be true
that:
G'{w)

G"{w)

(12-7)

at all frequencies. From Equations 2-65 and 2-66, it can be shown


that this is true only when the relaxation modulus is given by:

G{t) = Sr 1/ 2

(12-8)

where S is called the "strength of the network at the gel point."


Using this relaxation modulus equation in the Boltzmann superposition principle, Equations 2-7 or 2-8, it is predicted that 7] = 00
and G(oo) = O. Winter et al. [78,79] found that Equation 12-8 seems
to be valid for PDMS and polyurethane resins, but only when the
initial cross-linking agent concentration is at least adequate to carry
the curing process to the point where there is a maximum in the
modulus. For lower concentrations, it was found that Equation 12-7
was not satisfied but that G'(w) and G"(w), while not equal, were
both proportional to w n . This implies that the loss angle is independent of frequency and equal to WTr /2, where 1 > n > 0.5. This
implies that the relaxation modulus is given by:

G{t)

St- n

(12-9)

We note that the use of such a criterion requires that data be


obtained over a broad range of frequencies. When the amount of
cross-linking agent is just sufficient to reach gelation, a theoretical
analysis [80] suggests that n = 0.7.

CHEMORHEOLOGY OF REACTING SYSTEMS

421

Once the gel is formed, rubber elasticity theory can be used to


relate the modulus to the cross-link density [23,65,81].
REFERENCES
1. A. V. Tobolsky, I. B. Prettyman and J. H. Dillon, J. Appl. Phys. 15:324 (1944).
2. R. D. Andrews, A. V. Tobolsky and E. E. Hauson, J. Appl. Phys. 17:352
(1946).
3. K Murakami and K Ono, Chemorheology of Polymers, Elsevier Scientific
Publishing Co., New York, 1979.
4. Z. Kemblowski and J. Torzecki, Rheol. Acta 22:34 (1983).
5. C. A. May, Editor, "Chemorheology of Thermosetting Polymers", ACS Symposium Series, No. 227 (1983).
6. c. A. May, M. R. Dusi, J. S. Fritzen, D. K Hadad, M. G. Maximovich, K G.
Thrasher and A. Wereta, Jr., ACS Symp. Series, No. 227, p. 1 (1983).
7. R. J. Hinrichs, ACS Symp. Series, No. 227, p. 187 (1983).
8. D. M. Hoffman, ACS Symp. Series, No. 227, p. 169 (1983).
9. A. V. Tungare, G. C. Martin and J. T. Gotro, Polym. Eng. Sci. 28:1071 (1988).
10. M. B. Roller, Polym. Eng. Sci. 26:432 (1986).
11. s. P. Sawan, K Muni and J. Figlan, SPE Tech. Papers 35:1678 (1989).
12. C. D. Han and K-W. Lem, ACS Symp. Series, No. 227, p. 201 (1983).
13. s. J. Perry, J. M. Castro and C. W. Macosko, J. Rheol. 29:19 (1985).
14. E. B. Richter and C. W. Macosko, Polym. Eng. Sci. 20:921 (1980).
15. J. W. Blake, W. P. Yang, R. D. Anderson and C. W. Macosko, Polym. Eng.
Sci. 27:1236 (1987).
16. C. W. Macosko, Fundamentals of Reaction Injection Molding, Hanser, Munich,
1989.
17. L. T. Manzione, Polym. Eng. Sci. 21:1234 (1981).
18. J. M. Castro and C. W. Macosko, A.I.Ch.E.J. 28:250 (1982).
19. R. R. Eley, J. Coatings Technol. 5, No. 718, 49 (Nov. 1984).
20. M. J. Hannon, D. Rhum and K F. Wissbrun, J. Coatings Technol. 48: No. 621,
p. 42 (1976).
21. Y. Otsubo, T. Amari, K Watanabe and T. Nakamichi, J. Rheo!. 31:251 (1987).
22. s. E. Orchard, Appl. Sci. Res. A 11:451 (1962).
23. P. J. Flory, Principles of Polymer Chemistry, Chapter 11, Cornell Univ. Press,
Ithaca, NY, 1953.
24. J. Enns and J. K Gillham, J. Appl. Polym. Sci. 28:2567 (1983).
25. G. L. Hagnauer, P. J. Pearce, B. R. LaLiberte and M. E. Roylance, ACS
Symp. Series, No. 227, p. 25 (1983).
26. F. R. Volgstadt and C. L. Sieglaff, Polym. Eng. Sci. 14:143 (1974).
27. J. E. Hess, Modem Plastics, Nov. 1971, p. 60.
28. ASTM D3123-72, "Standard Test Method for Spiral Flow of Low-Pressure
Thermosetting Molding Compounds," 1988 Annual Book of ASTM Standards,
Vol. 8.03, p. 4.
29. P. J. Heinle, SPE Tech. Papers: 25:426 (1979); 26:447 (1980).

422

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

30. A. Hale, M. N. Garcia, C. W. Macosko and L. T. Manzione, SPE Tech. Papers


35:796 (1989).
31. S. Tonogai and S. Seto, Polym. Eng. Sci. 21:301 (1981).
32. B. L. Lee, L. G. Pappas, V. L. Folt and C. E. Sitz, SPE Tech. Papers 35:1201
(1989).
33. M. T. Shaw, S. Burkert and D. W. Sundstrom, Rev. Sci. lnstrum. 49:1597
(1978).
34. S. C. Malguarnera, D. R. Carroll and M. A. Colaluca, lEe Prod. Res. Dev.
23:103 (1984).
35. J. M. Dealy, Rheometers for Molten Plastics, Van Nostrand Reinhold, New
York, 1982, Sect. 7.5.
36. D. W. Sundstrom and S. J. Burkert, Polym. Eng. Sci. 21:1108 (1981).
37. D. Harran and A. Laudouard, J. Appl. Polym. Sci. 32:6043 (1986).
38. M. R. Kamal and M. E. Ryan, Polym. Eng. Sci. 20:859 (1980).
39. ASTM D4473-85, "Standard practice for measuring the cure behavior of
thermosetting resins using dynamic mechanical procedures."
40. Y. Nabata, A. Mamada and H. Yamasaki, J. Appl. Polym. Sci. 35:155 (1988).
41. R. P. White, Jr., Polym. Eng. Sci. 14:50 (1974).
42. V. M. Gonzalez-Romero and C. W. Macosko, J. Rheol. 29:259 (1985).
43. I.-c. Choy and D. J. Plazek, J. Polym. Sci. B 24:1303 (1986).
44. J. A. Biesenberger, R. Kumar, R. Garritano and J. M. Starita, Polym. Eng.
Sci. 25:301 (1985).
45. J. A. Biesenberger and R. Kumar, Polym. Proc. Eng. 3:141 (1986).
46. D. Serrano and D. Harran, Polym. Eng. Sci. 29:531 (1989).
47. J. K. Gillham, Polym. Eng. Sci. 19:676 (1979).
48. C. Y. Yap and H. L. Williams, Polym. Eng. Sci. 22:254 (1982).
49. w. X. Zukas, W. J. MacKnight and N. S. Schneider, ACS Symp. Series No.
227, p. 223 (1983).
50. G. A. Senich, W. J. MacKnight and N. S. Schneider, Polym. Eng. Sci. 19:313
(1979).
51. I. J. Goldfarb, C. Y. C. Lee and C. C. Kuo, ACS Symp. Series, No. 227, 49
(1983).
52. c. Y. C. Lee, C. C. Kuo and I. J. Goldfarb, ACS Symp. Series, No. 227, 61
(1983).
53. c. Y. C. Lee, J. Appl. Polym. Sci. 27:407 (1982).
54. H. L. W. Chan and J. Unsworth, Eur. Polym. J. 21:377 (1985).
55. w. J. Mikols and J. C. Seferis, ACS Symp. Series, No. 227, p. 95 (1983).
56. S. A. Yaloff and W. J. Wrasidlo, J. Appl. Polym Sci. 16:2159 (1972).
57. P. Hedvig, Dielectric Spectroscopy of Polymers, John Wiley & Sons, NY (1975).
58. S. Senturia and S. Garverick, US Patent No. 4,423,371.
59. S. Senturia and N. Sheppard, Adv. Polym. Sci. 80:1 (1986).
60. J. Gotro and M. Yandrasits, Polym. Eng. Sci. 29:278 (1989).
61. J. W. Lane, R. K. Khatta and M. R. Dusi, Polym. Eng. Sci. 29:339 (1989).
62. D. Kranbuehl, S. Delos, M. Hoff, P. Hoverty, W. Freeman and J. Godfrey,
Polym. Eng. Sci. 29:285 (1989).

CHEMORHEOLOGY OF REACTING SYSTEMS

423

63. R. W. Wetton, G. M. Foster, R. D. L. Marsh, J. C. Duncan and M. M. J. Blow,


SPE Tech. Papers 35:1650 (1989).
64. M. R. Kamal, Polym. Eng. Sci. 14:231 (1974).
65. F. G. Mussatti and C. W. Macosko, Polym. Eng. Sci. 13:236 (1973).
66. M. B. Roller, Polym. Eng. Sci. 15:406 (1975).
67. J. D. Keenan, SAMPE Educ. Workshop, Sunnyvale, CA, 1980.
68. D.-S. Lee and C. D. Han, Polym. Eng. Sci. 27:955 (1987).
69. M. R. Dusi, C. A. May and J. C. Seferis, ACS Symp. Series, No. 227, 301
(1983).
70. c. J. Kojima, M. E. Hushower and V. L. Morris, SPE (ANTEC) Tech. Papers
32:344 (1986).
71. Y. A. Tajima and D. Crozier, Polym. Eng. Sci. 23:186 (1983).
72. T. H. Hou, SPE Tech. Papers 31:1253 (1985).
73. K. Horie, I. Mita and H. Kambe, J. Polym. Sci. A-J 8:2839 (1970).
74. S. D. Lipshitz and C. W. Macosko, Poly. Eng. Sci. 16:803 (1976).
75. A. Y. Malkin et al., Polymer 25:778 (1984).
76. C. Y. M. Tung and P. J. Dynes, J. Appl. Polym. Sci. 27:569 (1982).
77. H. H. Winter and F. Chambon, J. Rheol. 30:367 (1986).
78. F. Chambon and H. H. Winter, J. Rheol. 31:683 (1987).
79. H. H. Winter, P. Morganelli and F. Chambon, Macromolecules 21:532 (1988).
80. J. E. Martin, D. Adolf and J. P. Wilcoxon, Polym. Preprints 31, No.1, p. 83
(1989).
81. S. S. Labana, in Encyc. of Polym Sci. & Eng., pp. 350-395, John Wiley & Sons,
NY, 1986.

Chapter 13
Rheology of Thermotropic
Liquid Crystal Polymers

13.1 INTRODUCTION

Liquid crystallinity is a state of matter that is intermediate between


the crystalline solid state and the liquid state. Hence, it is sometimes called a "mesomorphic" state or "mesophase," and the
molecular structural elements responsible for its formation are
called "mesogens." Liquid crystals do not have the three dimensional order of position of the molecules of solid crystals and are
therefore capable of flowing like fluids. Unlike ordinary fluids,
however, they do possess some order over macroscopic distances.
This order may be solely that of orientation of the molecules, or it
may also involve some positional order. There are many combinations of types of order, resulting in a wide variety of types of liquid
crystal, which are described in a number of texts [1-3].
"Thermotropic nematic" polymers whose mesogens lie along the
backbones of the polymer chains are of interest to us because of
their processibility by conventional plastics processing techniques,
and because of their mechanical and physical properties. "Thermotropic" means that the liquid crystal state is achieved by heating the
material to a suitable temperature. Materials that become liquid
crystal by dissolving in a suitable solvent are called "lyotropic."
"Nematic" means that the order in the liquid crystal is due to
orientation of elongated segments of the molecules. It is intuitively
reasonable that dense packing of rigid rod-like molecules should
favor a parallel alignment, analogous to the packing of logs flowing
down a river. In fact, the existence of nematic order for systems of
424

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

425

rigid rods was predicted theoretically [4, 5] long before liquid crystal
polymers (LCPs) were first synthesised.
A polymer can have a rod-like shape for various reasons. The
first synthetic polymers found to develop liquid crystallinity were
polypeptides, which, in the appropriate solvents, become rod-like by
forming helices. Hydroxypropylcellulose and other cellulose derivatives that form lyotropic and also thermotropic mesophases may
also form helices. Great interest in LCPs was first excited, however,
by the cwnmercialization by duPont of aromatic polyamide
("aramid") fibers. These were made from poly-p-benzamide or
poly-p-phenylene terephthalamide. These molecules are very nearly
linear because different conformations cannot be achieved by the
low energy motion of rotation about a bond, but only by bond
stretching or bending, which require much more energy. They
therefore meet the theoretical requirements for the formation of
liquid crystals and this behavior has been observed experimentally
[6].
The combination of intrinsic molecular rigidity and of excellent
alignment of the molecules in fibers results in outstanding stiffness
and strength compared to other organic fibers. Because they have
much lower density than do inorganics and metals, aramid fibers
also have much higher specific strength and modulus [7]. The
problem with the aramids is that they melt at too high a temperature to be processed as thermoplastics. A search was therefore
initiated at a number of laboratories to look for lower melting
analogous structures that might yield beneficial effects on properties and processing.
An account of this search, and of the chemistry and properties of
the products, is given by Calundann and Jaffe [8]. It was found that
p-linked aromatic polyesters were suitable, but even these were too
high melting as homopolymers. The trick was to find the type and
amount of comonomer that would decrease the melting point
sufficiently without destroying the linearity and rigidity crucial for
liquid crystal formation. A number of such comonomer systems
were found. At this time products based on three of these are
available commercially [9]. Hoechst Celanese has a family of products based on p-hydroxybenzoic acid (HBA) copolymerized with
2,6-hydroxynaphthoic acid (Vectra); Amoco a series based on

426

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

HBA with p,p' -biphenol and terephthalic acid (TA), which may
include isophthalic acid (Xydar); and Eastman Kodak and licencees offer, as X7G and X7H, the combination first recognized as
a thermotropic polymer, based on HBA, TA, and ethylene glycol.
13.2 RHEOLOGY OF LOW MOLECULAR WEIGHT LIQUID CRYSTALS

There is a continuum theory, developed by Leslie and Ericksen


[1, 10, 11], that describes well the rheological behavior of smallmolecule nematics. Doi [12,13] has developed a molecular theory
for rod-like polymers, and this reduces to the Leslie-Ericksen (L-E)
theory in the limit of low shear rates. The Doi theory seems to be at
least qualitatively applicable to lyotropic LCPs [14,15]. However,
the LCPs show a variety of phenomena not observed in low molecular weight liquid crystals, and it is not clear whether these phenomena are predictable by available theories [16]. The rheology of the
thermotropic polymers of interest to us is not describable in many
cases by these theories. Nevertheless, a brief description of the
rheology of low molecular weight nematics is justified, because it is
the basis for the attempts to date to describe the more complex
rheology of the LCPs.
At rest the lowest energy state of a nematic is that in which, on
the average, the long axes of the molecules are parallel. Their
orientation is described by a vector called the "director." At any
instant of time not all of the molecules are oriented in this direction. The orientation of individual molecules fluctuates. The degree
of orientation is described by an "order parameter" S, defined in
terms of the angle E> of a molecular axis with respect to the director
by

(13-1)
For a random distribution of orientations S is zero, and for a
perfect alignment it is unity.
The director with respect to which alignment occurs is not
specified in the theory. It is determined by external forces such as
those of magnetic or electric fields, temperature gradients, or
surface forces at the boundaries of the liquid crystal. As a result of
these forces the director may not be a constant. In fact, it is highly

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

~
--=-:: .:::-..::------ -- ---:::..:----~ ---==-:=.,
~

--- -

SPLAY

---=~

---- --...........
TWIST

427

....- ------:..=--w::,=-=--=.~

~----------=;---

BEND

Figure 13-1. Schematic of director cUlvature strains: splay, twist, and bend. Dashed lines are
parallel to director orientation. Note that for twist the directors are perpendicular to the
plane of the figure at bottom of sketch. Adapted from Ref. 3.

probable that the orientation direction will vary throughout the


volume of a sample. The spatial variation of the director is described in terms of three "director curvature strains," namely, splay,
twist, and bend, illustrated in Figure 13-1.
Energy is required to produce any of these curvature strains, and
the total energy W is given by

where n is the director. The K s are director curvature elastic


constants [1, Ch. 3]. For low molecular weight liquid crystals the Ki
values have magnitudes on the order of 10- 12 to 1O- 11 N (10- 7 to
10- 6 dynes). These low magnitudes mean that very little energy is
associated with the director curvatures and that these materials are
easily ordered by small forces.
The Leslie-Ericksen constitutive equation [10, 11] for the flow of
liquid crystals includes terms involving the director n and the
spatial variation of W, and requires five independent viscous constants, rather than the one required for isotropic fluids. The magnitude of the director is assumed to be constant in the L-E theory;
this is equivalent to assuming that the flow does not affect the order
parameter S. To solve the flow equations one must specify not only
the velocity at the boundaries of the apparatus, but also the
orientation of the director at the boundaries. Although the equations are very complex, they have been solved for a number of flows
that are viscometrically interesting. The predictions of the theory

428

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

can be summarized as follows:


1. At a point in the flow far from the influence of the boundary
orientation, a simple shear flow causes the director to lie at an
angle 0 0 to the flow direction such that
(13-3)
For 0 0 to be real, the two Leslie coefficients 3 and 2 must
have the same sign, which is not necessarily observed. If 0 0 is
not real, no single orientation direction is predicted. It is
interesting to note that 0 0 does not depend on the shear rate.
2. Near the boundaries there is a competition between the orienting effect of the shear flow, assuming there is one, and the
boundary orientation. The effect of the walls extends a distance d into the fluid:

(K/rrr )1/2

(13-4)

where K is a representative elastic constant (Equation 13-2), y


is the shear rate, and 1] the viscosity of the fluid.
3. The viscosity of the liquid crystal depends on the orientation
of the director relative to the flow and gradient directions.
Advantage can be taken of this to measure the Leslie coefficients, by measuring the viscosity in situations where the
director is maintained in different desired directions by imposing a strong magnetic field. In the absence of an external field,
the director orientation is determined by competition between
the orienting tendencies of the boundaries and the velocity
gradient. As a result, the viscosity will depend on the capillary
diameter in Poiseuille flow, becoming independent of the
diameter when it is large compared to the distance d given by
Equation 13-4. The L-E theory predicts the experimental
observation that data from different capillaries will superimpose if the viscosity is plotted against the fluid velocity times
the diameter; this contrasts with the usual scaling for viscous
flow in which superposition is achieved by plotting against
shear rate, which is proportional to velocity divided by
diameter.

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

429

For comparison with the very few data that have been obtained
on polymeric liquid crystals it is useful to list the values of the
elastic and viscous constants for a typical small-molecule liquid
crystal. For p-azoxyanisole, the elastic constants are 0.7, 0.4, and
1.7 X lO-llN for splay, twist, and bend, respectively [17]. The
Leslie coefficients for N-{p-methoxy-benzylidene)-p-butylaniline

~,
dv.
dy

7Ja

V.
n

L
dv.

7Jb

V.

Figure 13-2. Geometric arrangements for measurement of "Miesowicz viscosities." Velocity


Vx is in x-direction, velocity gradient dvx/dy in ydirection. Director n is fixed by a strong
magnetic field: in z-direction for 'ria' in x-direction for 'rib' and in y-direction for 'ric' Order
of viscosities is 'ric > 'ria > 'rib'

430

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

(usually abbreviated MBBA) are [17]


a 1 =

6.5 cp;

a4

83.2 cp; as

-77.5 cp;

a 2 =

a3 =

-1.2 cp;

46.3 cp

Viscosities for specific director orientations were measured by


Miesowicz [1, p. 164], and are conventionally named after him. The
Miesowicz viscosities of MBBA with the director in the neutral
direction (71), in the velocity direction (7Jb)' and in the gradient
direction (7JJ are 41.6, 24.8, and 103.5 centipoise, respectively. The
geometrical arrangement is illustrated in Figure 13-2.
Before we move on to consider the rheology of LCPs, it is
necessary to discuss one other phenomenon. The theory discussed
so far requires that the director orientation vary continuously. In
practice, however, unless special precautions are taken, a liquid
crystal will contain many points or lines at which there are discontinuities of the orientation. These defects, called "disclinations,"
have been classified, and a number are illustrated in Figure 13-3.
For small-molecule liquid crystals it is possible to remove these
defects and to maintain flowing samples free of them at moderate

Figure 13-3. Examples of defects ("disclinations") in nematic liquid crystals. Lines are
parallel to director. Points indicate singularities, where director orientation is indeterminate.
Adapted from Ref. 3.

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

431

flow conditions. As we shall see, this is not so with high molecular


weight thermotropic polymers. Furthermore, it is known that defects influence rheological behavior. Conversely, flow affects the
defect structure. We shall see, then, that the rheology of LCPs is
considerably more complex than that of low molecular weight liquid
crystals.
13.3 RHEOLOGY OF AROMATIC THERMOTROPIC POLYESTERS

The rigid, linear mainchain thermotropic polymers, of which the


aromatic polyesters are the most prominent examples, exhibit a
variety of rheological phenomena not found in homogeneous
isotropic polymers. An example of the variety that can be found in
one polymer merely by changing the test temperature is shown in
Figure 13-4. At 340C the flow curve is quite uninteresting, approaching a constant viscosity at low shear rates, with slight shear
thinning as the shear rate is increased. At 300C the flow curve has
the "Three Region" shape proposed by Onogi and Asada [19] as
characteristic of all LCPs. In their classification "Region I" is the
low shear rate region of shear thinning behavior. "Region II" is the
plateau of nearly constant viscosity at intermediate shear rates, and
this is followed by another shear thinning range, which is designated "Region III."
As the temperature is decreased further, it appears as though
Regions I and III are merging, and the flow curve approaches
power law behavior. Near 320C the flow curve is most peculiar,
with a range of shear thickening behavior. This has been observed
for a number of LCPs using both capillary and rotational instruments. It is probably not an artifact, therefore, and represents one
of a number of as yet unexplained anomalies in the behavior of
LCPs.
On the basis of rheo-optical measurements, Onogi and Asada
[19] concluded that the low shear rate range of shear thinning
behavior ("Region I") is one in which flow occurs by relative
motion of a dense pile of "domains." The structure of the domains
was not specified in detail, but it was postulated that a domain
consists of volume elements whose directors vary in such a way that
there is no net orientation when averaged over the volume of the
domain. In "Region II" the domains were considered to be trans-

432

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


105~--------------------------------------~

10 4

280C
290C
300C

- - - 340C

10~

________

__________

10

__________

100

~~

1000

SHEAR RATE (s-')

Figure 13-4. Variation of flow curve shape with temperature for a rigid thermotropic
polyester. Adapted from Ref. 18. Copyright 1980 by Elsevier Applied Science Publishers
Ltd. Reprinted by permission.

formed by increasing shear stress into a nematic fluid that acted as


a suspending medium for the remaining domains. Only in "Region
III" was the structure considered to be that of an oriented nematic
fluid.
The domain concept was turned into a quantitative theory for the
shear rate dependence of the viscosity by Marrucci [20]. He considered a domain to be a volume element surrounding some cluster of
disclinations, such as is illustrated in Figure 13-Sb. Although the
director strain energy of this arrangement is higher than that of the
uniform orientation state shown in Figure 13-Sa, the arrangement is
in .a local energy minimum. There is no path by which it can achieve

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

(a)

433

(b)

Figure 13-5. Schematic of Marrucci's model of a domain (b) as a cluster of disclinations,


compared to lowest energy state (a) with uniform director orientation. Adapted from Ref. 20.

the lower energy state of uniform orientation without going through


much higher energy states. If the energy barrier is greater than the
thermal energy kT, the domain can be considered to be a stable
entity. As in Onogi and Asada's model, a domain is considered to
have no net orientation.
Marrucci used the domain model and the concepts of director
curvature strain and of anisotropic viscosity of nematic fluids to
predict the existence of Regions I and II of the flow curve. The
transition from Region I to II was consistent with the measured
magnitude of the elastic constants and with microscopic observations of domain size. Wissbrun [21] subsequently used Marrucci's
picture of a domain in a somewhat different model having a greater
flexibility in the prediction of the Region I slope, although at the
cost of adding another parameter.
The domain models were intended to account for the surprising
low shear rate behavior of LCPs, which is characterized by shear
thinning viscosity and by absence of orientation of the director in
shear flow. Before proceeding, it is appropriate to consider possible
reasons for the importance of domains in LCPs, especially the
thermotropic polymers of interest here. Also, it is necessary to
caution the reader that there may be other reasons, having no
connection with liquid crystallinity, for what appears to be Region I
flow behavior.
It appears that the density of a domain texture, the time and
effort required to eliminate it, and its effect on the rheological
behavior, all increase as an LCP becomes stiffer or increases in
molecular weight, or becomes more concentrated, in the case of
lyotropic LCPs. It is difficult to find systematic and quantitative

434

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

data to substantiate this generalization, but it seems consistent with


the reported observations. To some extent this could be a kinetic
effect, because the viscosity also increases with molecular weight
and stiffness. In addition, it is likely that LCPs are inherently more
likely to form disclinations than low molecular weight liquid crystals, because they have substantially higher anisotropy of the elastic
constants and of the Leslie viscosity coefficients. These have not
been measured on thermotropic polymers, but data on lyotropics
[16] are suggestive. The elastic constants for bend and splay are 7 to
16 times higher than for twist in two such LCPs. For the low
molecular weight liquid crystal discussed above, the maximum ratio
of these constants was 4. Similarly, the ratio of viscosity coefficients
for bend and twist was as small as 0.006, compared to 0.24 for the
MBBA also discussed above.
The other cause for the appearance of Region I behavior arises
from the fact that the thermotropic polymers of interest are all
copolymers. The homopolymers from each of the constituent
comonomers have very high melting points. Relatively short sequences of these monomers may form small and imperfect crystals
with melting points higher than the nominal melting temperature of
the copolymer. These crystallites act as crosslinks for the copolymer
melt and change the rheological behavior drastically from what it
would be in their absence.
A striking example of this effect is illustrated in Figure 13-6 [18].
The three flow curves in this figure were all measured at 210C. The
difference among them is that the polymer was loaded into the
rheometer at different temperatures, one in the vicinity of and one
well above a DSC endotherm. The melts were then cooled to the
measuring temperature. Dilute solution viscosity measurements of
the extrudates showed that the striking differences in the shapes of
the flow curves did not result from degradation. In addition, separate viscometry experiments showed that the effect was reversible
with time, and DSC and x-ray diffraction analyses supported the
hypothesis that the thermal history effect was caused by the melting
of crystallites.
It should be noted that the phenomenon is by no means unique
to LCPs; a similar explanation has been given for the extreme
temperature dependence of the melt rheology of polyvinyl chloride
[22]. Crystallites may also grow when the melt is held in an appro-

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

435

103r-----~----------------------------~

10~

100

________

__________

10'

__________

10 2

~~

10 3

SHEAR RATE (s-')

Figure 13-6. Effect of thermal history on flow curve of a crystallizable thermotropic LCP. All
flow curves were measured at 210C by a capillary viscometer with a die having an LjR of
59, but loaded at various temperatures and then cooled to 210C. Loading temperatures,
from top to bottom: 210, 240, 300C. Adapted from Ref. 18. Copyright 1980 by Elsevier
Applied Science Publishers Ltd. Reprinted by permission.

priate temperature range, and this will cause the viscosity to increase with time [23]. This effect has obvious consequences both for
laboratory measurement and for processing operations in which
long melt residence times may occur.
The reason for the emphasis on Region I flow behavior is that
the rheology of the rigid thermotropic polymers that have been
commercialized is of this type over a wide range of shear rates and
temperatures. An example is shown in Figure 13-7, where we see
power-law flow over the entire range of measurement shown. Only
at much higher shear rates, above 10 4 s -1, does the flow curve show
evidence of a plateau region [23]. The only anomaly in Figure 13-7
is the strong temperature dependence of the viscosity in the region
of 280 to 300C, This is probably caused by the melting of small
crystallites in this temperature range.
The evidence that this power-law flow region is in fact what we
have called Region I flow is that little or no orientation is developed in shear flow over the shear rate range shown. Strong orientation is caused, however, by extensional flow at strains modest
compared to those required to orient flexible polymers [24].

436

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

104~---------------------------------------------------,

10~

10

________________ ______________
~

10 2

~~

10 3

______________

10 4

SHEAR RATE (5-')

Figure 13-7. Temperature dependence of flow curves of a copolyester of p-hydroxybenzoic


acid and 2,6-hydroxynaphthoic acid measured by capillary viscometry, with die LjR of 59.
Temperatures, from top to bottom, are: 280, 290, 300, 310 and 350C. Adapted from Ref. 23.
Copyright 1987 by Gordon and Breach Science Publishers. Reprinted by permission.

It should be noted that the magnitude of the viscosities in the


shear rate range around 10 3 s -1 is quite low compared to that of
isotropic polymers of comparable molecular weight. For instance,
such a polyethylene terephthalate has a zero shear rate viscosity of
about 3000 poise, and is just beginning to shear thin at 10 3 s -1 [8].
The rheological behavior of LCPs differs from that of flexible
polymers in many other respects. They have an unusually large end
correction for capillary flow [23], a phenomenon associated with
high melt elasticity and/or high extensional viscosity when observed
in flexible polymers. Consistent with that behavior, they display a
higher first normal stress difference than do flexible polymers [25].
Quite inconsistently, however, they show very little extrudate

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

437

swell [18,25]. Another anomaly is that the first normal stress


difference N1 is about 10 times higher than the dynamic storage
modulus G', compared with the factor of about 2 usually observed
for isotropic polymers. The Cox-Merz rule, i.e., equality of the
steady shear and dynamic viscosities, mayor may not be obeyed,
depending on test conditions [18]. The transient behavior on start-up
of steady shear is unusual as well. Large maxima of shear stress are
observed at strains on the order of 50 strain units, independent of
shear rate [18,26]. Shear thickening is another anomaly that was
already shown in Figure 13-4 for capillary flow and that has also
been observed for another thermotropic polyester in a cone and
plate rheometer [27].
There is also a body of results that suggests that the domain
texture that can be observed microscopically is sensitive to shear
history and that a change of orientability and of rheology accompanies the change of texture [27]. Last in this long list of peculiar
forms of flow behavior is the observation of a negative N 1, which
was first observed in lyotropic LCPs, but subsequently also in
thermotropics [28]. It has been speculated that the negative N1 is
associated with bands of orientation that are often observed upon
relaxation of shear of LCPs. The bands are transverse to the shear
direction, and have alternating plus and minus 45 orientation
angles.
It is clear that we are a long way from understanding and being
able to explain all of these complex and unusual observations.
Nevertheless, we shall see in the next section that the observations
themselves can be used to make useful statements about the processing of these polymers.
13.4 RELATION OF RHEOLOGY TO PROCESSING OF LCPs

The rheological and physical properties that are important to the


processing of thermotropic LCPs can be summarized as follows:
1. Shear thinning, leading to low viscosity at high shear rates and
high viscosity at low shear rates.
2. Low extrudate swell.
3. Little orientation in shear, strong orientation in extension.

438

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

4. High activation energy for viscosity, and very low heat of


fusion.
5. Low (anisotropic) coefficients of thermal expansion.
The roles played by these properties in processing, particularly in
injection molding are:
1. The low viscosity at high shear rates makes it possible to fill
large molds [27] and molds with long, thin-walled sections [29].
The high viscosity at low shear rates provides good sag resistance for extrusion applications.
2. The low extrudate swell requires gates of injection molds to be
larger than is normally required for isotropic melts, in order to
avoid "jetting," which produces multiple weld lines in a part.
The low extrudate swell facilitates die design for extrusion of
precise profiles.
3. The thermal properties of the melt are responsible for rapid
solidification in a mold and lead to short cycle times. They will
also, however, slow the knitting of weld lines. Molds and
processing conditions must be designed, therefore, to insure
that weld lines, if any, occur in non-critical regions of the part.
4. Because the flow field in mold filling is complex, the difference
in the orientation resulting from shear and extension gives rise
to a complex morphology [24]. Such behavior, including the
formation of multiple layers with various degrees and directions of orientation, are also observed with short fiber filled
conventional thermoplastics [30]. The mechanical response of
the molded part is also quite analogous for LCPs and fiber
reinforced isotropic polymers [24]. The variability of orientation, and therefore of mechanical properties, is a potential
pitfall but also offers the opportunity to tailor the final part
properties to a given application.
5. The low coefficient of thermal expansion and high rigidity
minimize warpage and favor high precision molding. The
anisotropy of the coefficient of expansion results from the
orientation, and, if properly controlled, can be used to advantage. An example is the ability to match the coefficient of
expansion of glass for jacketed optical fiber applications [31].

RHEOLOGY OF THERMOTROPIC LIQUID CRYSTAL POLYMERS

439

The above summary illustrates the applicability of the rheological


and other physical measurements to the prediction of processing
behavior, despite the complexity of the rheology of these materials.

REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.

P. G. deGennes, The Physics of Liquid Crystals, Oxford University Press,


Oxford, 1974.
H. Kelker and R. Hatz, Handbook of Liquid Crystals, Verlag Chemie, Deerfield Beach, Florida, 1980.
D. Demus and L. Richter, Textures of Liquid Crystals, Verlag Chemie, New
York,1978.
L. Onsager, Ann. N.Y. Acad. Sci. 51:627 (1949).
P. J. Flory, Proc. Roy. Soc. A 234:73 (1956).
S. L. Kwolek, P. W. Morgan, J. R. Schaefgen and L. W. Gulrich, Macromolecules 10:1390 (1977).
J. P. Riggs, "Carbon Fibers," pp. 640-684, in Encyclopedia of Polymer Science
and Engineering, Second Edition, Vol. 2, Wiley, New York, 1985.
G. W. Calundann and M. Jaffe, Proc. Robert A. Welch Conferences on
Chemical Research XXVI. Synthetic Polymers, Houston, Texas, Nov. 15-17,
1982.
C. E. McChesney, "Liquid Crystal Polymers," in Engineered Materials Handbook, Vol. 2, ASM International, Metals Park, Ohio, 1988.
J. L. Ericksen, "The Mechanics of Nematic Liquid Crystals," in The Mechanics of VIScoelastic Fluids, ed. R. S. Rivlin, AMD Vol. 22, ASME, New York,
1977.
F. M. Leslie, "Theory of Flow Phenomena in Liquid Crystals," in Advances
in Liquid Crystals, Vol. 4, ed. G. H. Brown, Academic Press, New York, 1979.
M. Doi, J. Polym. Sci., Polym. Phys. Ed. 19:229 (1981).
N. Kuzuu and M. Doi, J. Phys. Soc. Japan 52:3486 (1983); 53:1031 (1984).
G. C. Berry, Mol. Cryst. Liq. Cryst. 165:333 (1988).
A. B. Metzner and G. M. Prilutski, J. Rheol. 30:661 (1986).
K. F. Wissbrun, J. Rheo!. 25:619 (1981).
R. G. Larson, Constitutive Equations for Polymer Melts and Solutions, Chapter
10, Butterworths, Boston, 1988.
K. F. Wissbrun, Br. Polym. J., 12:163 (1980).
S. Onogi and T. Asada, "Rheology and Rheo-optics of Polymer Liquid
Crystals," pp. 127-147 in Rheology, Vol. I, G. Astarita, G. Marrucci, and L.
Nicolais, eds., Plenum, New York, 1980.
G. Marrucci, Advances in Rheology, Vol. 1,441-448, B. Mena, A. Garcia-Rejon, and C. Rangel-Nafaile, eds., Universidad Nacional Autonoma de Mexico,
1984.

440

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

21.
22.

K. F. Wissbrun, Faraday Disc. Chem. Soc. 79:161 (1985).


L. A. Utracki, Z. Bakerdjian and M. R. Kamal, Trans. Soc. Rheol. 19:173

23.
24.
25.
26.
27.
28.
29a.
29b.
30.
31.

(1975).
K. F. Wissbrun, G. Kiss and F. N. Cogswell, Chem. Eng. Commun. 53:149
(1987).
Y. Ide and Z. Ophir, Polym. Eng. Sci. 23:261 & 792 (1983).
A. D. Gotsis and D. G. Baird, J. Rheol. 29:539 (1985).
G. G. Viola and D. G. Baird, J. Rheol. 30:601 (1986).
F. N. Cogswell, "Observations on the Rheology of Thermotropic Polymer
Liquid Crystals," Ch. 10 in Recent Advances in Liquid Crystalline Polymers,
L. L. Chapoy, ed., Elsevier, New York, 1985.
M. Prasadarao, E. M. Pearce and C. D. Han, J. Appl. Polym. Sci. 27:1343
(1982).
G. E. Williams, "Thermotropic Liquid Crystal Polymers," in Special Polymers
for Electronics and Opto-Electronics, M. Goosey, ed., Elsevier, New York,
1989.
C. E. McChesney and J. R. Dole, Modem Plastics, January 1988.
D. McNally, Polym.-Plast. Techno!. Eng. 8:101 (1977).
F. Yamamoto, Mol. Cryst. Liq. Cryst. 153:423 (1987).

Chapter 14
Role of Rheology
in Extrusion
14.1 INTRODUCTION

Extrusion is the most important single polymer processing operation. Virtually every pound of thermoplastic polymer is subjected to
an extrusion process at some point in its conversion to a finished
article. It is more amenable to theoretical analysis than some other
processing operations for a number of reasons:
1. It is a continuous, steady state process, not discontinuous like

injection molding,
2. For the most common mode of operation there are no free
surfaces within the extruder, so that boundary conditions can
be prescribed on known surfaces, and
3. Viscoelastic behavior plays only a minor role, and viscous fluid
models have been found adequate for the analysis.
The technological significance of extrusion and the possibility of
analysis have motivated a considerable body of research and a
correspondingly advanced state of understanding of the process. In
this chapter we present an overview of the state of this subject with
an emphasis on the role played by melt rheology.
A single chapter cannot hope to cover the topic of extrusion in
sufficient detail for all needs. It is appropriate, therefore, to give
some general references for more detailed expositions. A very
comprehensive reference is "Polymer Extrusion" by Rauwendaal
[1]. It covers not only the analysis literature but also information on
hardware, instrumentation, and practical operation. Stevens [2]
441

442

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

gives a useful 20 page table on operational strategies. Other books


are listed in the references at the end of this chapter [3-6].
14.1.1 Functions of Extruders

For our purpose an "extruder" is a machine that, by the action of a


screw, converts a solid polymer to a melt and generates pressure to
force it through a shaping die. There are situations, usually in
polymerization or compounding, in which the feed to the extruder
is already molten; in this situation the extruder need not carry out
the melting step, which is sometimes called "plastication." Also, for
some processes such as fiber spinning, a screw extruder is not a
sufficiently precise pumping device. The function of the extruder is
then merely to generate the required pressure, and the metering is
done by a gear pump fed by the extruder.
However, in the great majority of applications the feed to the
extruder is a solid and the extrudate goes directly to a shaping
operation. The extruder is required to compact and melt the solid,
generate the pressure needed, and meter the flow of the molten
polymer through the shaping die at the desired rate and temperature.
There are also other tasks that extruders are required to perform.
They are used, for example, to blend miscible polymers to form a
homogeneous mixture. These might be chemically identical but
have different molecular weights. Or, they may be essentially identical but differ in physical form. For instance, in many processes
regrind or scrap material is mixed with virgin polymer, and it is
necessary to achieve good dispersion at least to average out the
possible effects of prior processing history.
Blending of chemically dissimilar polymers is another process
often done by extrusion. In some instances the polymers are miscible. In this case the requirement is to produce a mixture that is
homogeneous in composition. More likely they are immiscible, in
which case it is necessary to produce not only macroscopic homogeneity but also to achieve some desired morphology and specified
disperse phase size distribution.
Mixing is important in extrusion for another reason. Heat is
generated during the extrusion process, generally in a non-uniform
fashion, so that some of the molten polymer is heated to a higher

ROLE OF RHEOLOGY IN EXTRUSION

443

than average temperature. Continuous mixing of the portions of the


melt at different temperatures is desirable to avoid overheating and
thus degradation or discoloration. Also, melt viscosity varies with
temperature, and because the viscosity affects the rate of extrusion
and may affect the behavior in downstream processing, in order to
achieve a uniform product it is necessary that the melt exiting from
the extruder have maximum temperature homogeneity.
Another function that an extruder may have to perform is devolatilization, the removal of volatile components from the feed.
The volatile material might simply be air adsorbed on a powdery
feed or it could be residual monomer or solvent from the polymerization process. Volatiles can also be generated during extrusion as
a product of a chemical reaction being conducted in the extruder.
Normally one thinks of an extruder as a machine operating in a
continuous fashion. However, the same basic machinery is also used
as the means of plasticating solid polymers in discontinuous processes such as injection molding and some types of blow
molding.
14.1.2 Types of Extruders

A classification [1, p. 24] of extruder types is shown in Table 14-1.


Single screw extruders are the most commonly used. They are sold
in a large range of sizes, from 3/4 inch inside barrel diameter, with
an output of a few pounds per hour, to 20 inch diameter, capable of

Table 14-1. Types of Extruders


I.

II.

III.
IV.

DISCONTINUOUS, RECIPROCATING
Ram Extruders
DISCONTINUOUS, RECIPROCATING/ROTATING
Reciprocating Screw Extruders
CONTINUOUS, RECIPROCATING/ROTATING
Buss-Kneader, etc . .
CONTINUOUS, ROTATING
A. Screw Extruders
1. Single screw
2. Twin Screw-Counterrotating or Corotating
Self-wiping, partial intermeshing,
or tangential
B. Disk or Drum Extruders

444

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

extruding on the order of 20,000 pounds per hour. The length-todiameter (LID) ratio can be very small for melt-fed extruders used
only as pumps; more commonly the LID is on the order of 20 to 30
for plasticating extruders, and may be even longer for extruders
fitted with vent ports for devolatilization. Further, there is a bewildering variety of screw designs to achieve special objectives, such as
enhanced melting rate or mixing or the addition of reinforcing
fibers to molten polymer.
Single screw extruders, although widely diverse in design, do have
certain inherent limitations. In response to these, various types of
twin screw extruders have been developed. These are becoming
increasingly popular for certain applications, despite their higher
capital cost.
A number of extruder designs have been proposed that do not
use the screw principle at all. Rather, they use other rotating
elements such as disks or drums. We do not consider them further
in this book; the interested reader is referred to Rauwendaal [1] for
descriptions of some of them.
14.1.3 Screw Extruder Zones

Figure 14-1 shows a schematic of a simple single-screw plasticating


extruder. In this diagram the screw has a single helical flight, with a
constant helix angle. The barrel is cylindrical, and the outside
diameter of the screw flights is constant. The diameter of the root
of the screw, and therefore the depth of the screw channel, varies
along the length of the extruder. At the feed end of the extruder,
where solid polymer is admitted, the channel is relatively deep. At
the output end, where the melt forming die is attached, the channel
is much more shallow. Connecting these two zones is a transition
section with a tapering channel depth.
These three geometrically different zones correspond to the basic
three tasks that a plasticating extruder must perform:
1. To compact a loose or granular feed into a dense solid form,
2. To convert the solid to a melt,
3. To convey the molten polymer and to exert pressure to force it
through the die that shapes it into the desired form.

ROLE OF RHEOLOGY IN EXTRUSION

445

HOPPER

Figure 14-1. Schematic of single-screw plasticating extruder. Adapted from Ref. 30. Copyright 1983 by Hanser Publications, Munich, Vienna, NY; Distributed in USA and Canada
by Oxford University Press. Reprinted by permission.

This summary does not mean that each of the above tasks occurs
only in the designated zones shown in the schematic. On the
contrary, melting begins before the tapered transition zone, and
pressure development may occur all along the screw. On the whole,
however, the design of the screw geometry does reflect the three
basic functions of an extruder.
Among the other tasks of extruders mentioned above is devolatilization. To accomplish this a "multi-stage" screw is necessary.
In such a design there is a zone in which the pressure of the melt is
reduced to zero in order that the melt not be forced out of the vent
in the extruder barrel through which the volatiles are removed.
Similar decompression zones are required for the addition of components to a formulation in an extruder. It may be advantageous,
for example, to feed reinforcing fiber into an extruder at a point
where the polymer component is molten, thereby minimizing wear
on the extruder and breakage of the fiber. In either case, the
decompression zone must be followed by another pumping or
conveying zone to generate pressure.
Modifications of the basic design have been invented for specific
needs such as intensive mixing, increased melting capacity, and
minimization of melt temperature rise [1].

446

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

14.2 ANALYSIS OF SINGLE SCREW EXTRUDER OPERATION

The analysis of extruder operation relates operating conditions,


such as flow rate, pressure, and temperature, to polymer material
properties and to extruder design parameters. Such an analysis can
be used for screw design, to optimize operating conditions, or to
select or modify resin characteristics.
14.2.1 Approximate Analysis of Melt Conveying Zone

The melt conveying zone of a single screw extruder is the one that
has been analyzed most completely, and there is considerable
confidence that the results are generally applicable. By making
certain approximations it is possible to express the results in the
form of simple explicit equations, which are useful as first approximations of the operation of this zone. The results of more complete
analyses can then be expressed graphically or numerically in the
form of corrections to the simple equations. The simple model also
serves to illustrate some important concepts and problems in the
simulation of the entire extruder.
The main approximations of the simple model are that:
1. The polymer melt is a Newtonian fluid whose viscosity is

independent of temperature,
2. The depth of the channel in which the melt flows between the
screw and the barrel is small compared to its width and to the
diameter of the extruder, but is large compared to the clearance between the screw flight and the barrel.
The geometric approximations are relatively mild, and the effects
of removing them can be estimated as numerical correction factors.
The effects of assuming the melt viscosity to be independent of
shear rate and temperature are much more serious and more
difficult to analyze. In general, as is described in Section 14.2.3, the
more exact analyses replace this approximation by that of a power
law fluid with an exponential temperature dependence.
Figure 14-2a is a schematic of a short section of a single-flighted
screw. The geometric approximations allow the replacement of this
complex geometry by the simple one shown in Figure 14-2b. The

ROLE OF RHEOLOGY IN EXTRUSION

(a)

447

d,

FLIGHT

~BARREL~~~

T
1

ROOT

(b)

BARREL

T
ROOT OF THE SCREW

Figure 14-2. Schematic of screw and of its geometric approximation: (a) Section of screw and
definitions of symbols. Adapted from Ref. 8. Copyright 1977 by McGraw-Hill Publications.
Reprinted by permission. (b) Simplifying approximations to screw geometry made in analysis
of melt conveying zone in Section 14.2.1. Adapted from Ref. 3.

approximation amounts to considering that the channel had been


formed by wrapping a "tape" with a rectangular cross-section as a
helix along a cylindrical rod. The thickness of the tape corresponds
to the channel depth H, the width to the perpendicular distance W
between flights, and the diameter of the rod to the diameter D of
the root of the screw. The analysis then considers the motion in the
tape as though it had been unwrapped from the rod.

448

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The surfaces of the root of the screw and of the barrel are now
parallel planes. The channel is bounded by plane flight surfaces.
For convenience of mathematical analysis it has become conventional to consider the barrel to be moving with respect to a
stationary screw. The linear velocity Vb of the barrel is inclined at
the helix angle (J to the down-channel direction z. The transverse
or cross-channel direction is denoted as the x coordinate, and the
radial direction as the y coordinate.
In terms of the actual extruder geometry and operating conditions, Vb is given by
(14-1)
where N is the angular screw speed (in revolutions per unit time).
There are two driving forces for the flow of the melt: drag flow
and pressure flow, as in some of the viscometric flows described in
Chapter 4. Drag flow occurs when the melt adheres to two solid
surfaces, one of which moves relative to the other. Thus, for
example, the moving surface in Figure 14-2b generates drag flow,
with the flow velocities in planes parallel to the surface. The drag
flow velocity profile can be expressed in terms of a down-channel
component V z and a transverse component vX ' each of which
depends upon the vertical distance y from the bottom surface.
There is also back flow because of pressure gradients that result
from the build up of pressure at the end of the extruder due to the
resistance to flow offered by the die. A pressure gradient can also
result from melt being forced into the zone under pressure. The
pressure flow between the parallel surfaces is also parallel to these
planes. Therefore we can assume that the Vy component of velocity,
perpendicular to the bounding planes, is zero everywhere except in
the immediate vicinity of the screw flights.
So far we have assumed that the channel is wide, and the flow
problem has been treated by completely ignoring the existence of
the screw flights. This assumption is justified over most of the width
of the channel. However, the channel is not infinitely wide and the
screw flights do exist. At some time material flowing in the x-direction will encounter the impenetrable flight and be forced to reverse
its flow. Physically there is no net flow in the cross-channel direction. Whatever flow occurs at one plane in the channel must be

ROLE OF RHEOLOGY IN EXTRUSION

449

balanced by flow in the opposite direction at some other plane. (We


have neglected the "leakage flow" between the screw flights and
the barrel.) The cross-channel velocity profile calculated with the
above assumptions is
(14-2)

where Vbx (= Vb sin 0) is the cross-channel component of the relative velocity of the bounding surfaces given by Equation 14-1.
Equation 14-2 is a parabolic function of y. The cross-channel
velocity is zero at the screw root (y = 0) and at a height two-thirds
of the way up the channel (y = 2H/3). At the barrel surface
(y = H) it is, as required by the no-slip boundary condition, equal
to vbx ' the transverse component of the barrel velocity. The complete velocity profile is shown graphically in Figure 14-3. Although
Vx does not contribute to the output of the extruder, it is an
important parameter for the following reasons:
1. The cross-channel flow consumes power and must be known
for sizing the extruder motor and for calculating the torque
exerted on the screw.
2. The power consumed is dissipated as heat, which raises the
melt temperature if it is not removed by convective and
conductive heat transfer. Temperature rises will affect the
extrusion operation by changing the viscosity and must be
controlled to prevent degradation.
3. The circulatory cross-channel flow is the primary mechanism
for mixing and homogenizing the melt in a single-screw extruder.
4. For a shear-thinning melt, shear associated with the crosschannel flow reduces the viscosity and thereby changes the
down-channel velocity profile and the extruder output.

The down-channel velocity profile can be expressed as the sum of


two components. One is that due to the drag of the moving plane.
The other is due to the down-channel pressure gradient. This
gradient is constant throughout the flow channel under the present
assumptions. The magnitude of the pressure gradient depends on

450

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

(a)

(e)

(b)

<1>=0

<1>=(1/3)

~
~
<I> =

(2/3)

~
~

1.0
0.8

0.6
<1>=1

0.4
0.2
0
-4 -2

vbx

2 -4 -2

8 10 0

~
Vbz

V
Vb

Figure 14-3. Velocity profiles in screw channel for several values of 1>, the ratio of pressure
to drag flow, from Equations 14-2 and 14-3. The ordinate is the relative distance from the
root of the screw to the barrel surface. Top (1) = 0) is open discharge condition; bottom
(1) = 1) is closed discharge. (a) Cross channel velocity, vr (b) Down channel velocity, v z .
(c) Axial velocity, v. Adapted from Ref. 9. Copyright 1962 by John Wiley & Sons, Inc.
Reprinted by permission.

factors external to the melt conveying section, such as the die flow
resistance, so we cannot specify it numerically as yet. The downchannel velocity is given by
1 JP
--y(H-y)

2YJ Jz

(14-3 )

ROLE OF RHEOLOGY IN EXTRUSION

451

where Vbz (= Vb COS 0) is the down-channel component of the


relative boundary plane velocity.
The first term of Equation 14-3 is the drag flow. The second term
is the parabolic pressure flow velocity profile for a Newtonian fluid.
The additivity of these two terms is not a general result and applies
only to Newtonian fluids. For a fluid whose viscosity depends upon
shear rate, such a separation of terms cannot be made. The minus
sign of the pressure flow term reflects the fact that a positive
pressure gradient (pressure increasing in the down-channel direction) opposes the down-channel drag flow.
It is useful to express Equation 14-3 in dimensionless form, using
the following variables:

Y-=y/H
<I>

H2 ap

-= - - 67JVbz az

In terms of these variables Equation 14-3 becomes

v = Y[1

- 3<1>{1 - Y)]

(14-4)

From the expression for the total flow given below it will be seen
that the group <I> is the ratio of the pressure flow to the drag flow.
For now it is a convenient parameter for describing how the velocity
profile varies with conditions. This is shown graphically in Figure
14-3. The left column shows the cross-channel velocity vX ' normalized by dividing by v bx The cross-channel velocity is independent
of the down-channel flow and is the same for all the cases illustrated. The normalized down-channel velocity V depends on the
parameter <I> according to Equation 14-4, and the velocity profile is
shown for various values of <I> in the center column of Figure 14-3.
In the third column the velocity along the barrel axis direction,
obtained from the vectorial addition of Vx and vz ' is shown. Consistent with the definition of <I> as a ratio of pressure to drag flow,
when <I> is zero the velocity profile of
is linear, as expected for
drag flow. When <I> = 1, the pressure flow just balances the opposing drag flow and there is no net flow in the barrel direction.

v:.

452

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The velocity profile of any element of the fluid is complicated. It


depends upon <I> and upon the initial location of the fluid element
in the channel. In general it can be described as a helical path
within the screw channel, which itself has the shape of a helix [3].
The flow rate (or "throughput") of the conveying section is
calculated by integrating vz over the width Wand height H of the
channel. The result is
Q=

vbzWH
2

WH 3 ap
--127]

az

(14-5)

Just as with the velocity profiles, the drag flow and the pressure
flow terms are additive, and it is readily verified that <I> is indeed
their ratio.
Having reached this result, we need to "rewind our tape" onto
the screw and replace the parameters of Equation 14-5 with those
of the actual extruder geometry. Recall that

and that
W = L cos 0 = 7TDb tan 0 cos 0 = 7TDb sin 0

Also, the pressure gradient ap /az is measured along the downchannel direction z. This is related to the pressure gradient ap /a/
measured in the axial length direction, /:

ap
az

ap
a/

-sinO

Making these substitutions into Equation 14-5 gives the output of


the melt conveying section as
7T2ND~H cos 0 sin 0
Q=------2

(14-6)

We will use this result in the next section, in which we analyze


the effects of the variables upon the performance of the extruder
when pumping melt through a die.

ROLE OF RHEOLOGY IN EXTRUSION

453

The final results of interest for the simplified model are the
power requirement and the viscous dissipation. The power E for a
length of channel Lis:

The first term of this equation is readily seen to be the power for
the down-channel drag flow. The second term arises from the
circulatory cross-channel flow, and the third term accounts for
the down-channel pressure flow. In the convention we are using,
the parameter <I> is positive when the pressure flow opposes the
drag flow; a positive <I> (or JP /Jz) therefore corresponds to a
positive contribution to the power required.
It is also of interest to use Equation 14-7 to determine the
efficiency of the screw extruder as a pump. In other words, we can
calculate the fraction of the power input that is used to generate
pressure to push the melt against the resistance of the die. The
remaining fraction is dissipated as heat. We will give the results of
this calculation but note that one of the simplifying assumptions we
have made is seriously in error with respect to the power requirement. That assumption is the neglect of the leakage flow over the
flight clearance. Because of the inverse dependence of the power
upon the depth H in Equation 14-7, the power associated with the
leakage flow is not negligible. We will return to this point in Section
14.2.3.2, but for the moment will proceed with the calculation. The
efficiency, e, is given by:
e=

3<1>( 1 - <I> )cos 2 0


4 - 3 cos 2 0(1 - <1

(14-8a)

or, alternatively, by:


e =

3<1>( 1 - <1
------~------

+ 4 tan 2 0 + 3<1>

(14-8b)

Solutions of this equation are given graphically by Booy [7] with


the helix angle 0 expressed in terms of L / D, recalling that it is

454

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

equal to 7T tan fJ. For a square pitch screw, with the length of the
lead equal to the diameter, the maximum efficiency is only about
28%; the highest maximum at any angle is 1/3.
Obviously the extruder is not a very efficient pump, with most of
the power input going into heat generation. It must be remembered
however, that in the melting zone of the extruder advantage is
taken of this "inefficiency" to generate heat to melt the polymer.
14.2.2 Coupling Melt Conveying to Die Flow

In this section the analysis of the melt conveying zone is used to


predict the performance of a system consisting of a melt conveying
zone and a shaping die. This is a useful exercise despite the number
of unrealistic assumptions and approximations that we have made
because:
1. The conclusions are applicable qualitatively to more realistic

situations,
2. The analysis provides a baseline for estimating the effects of
more realistic conditions, such as non-Newtonian and non-isothermal flow,
3. The analysis is similar in form to the one required to solve the
much more complicated equations for the general case of a
plasticating extruder, with its different zones that are coupled
together. It gives, then, a basis for understanding the principle
of operation of the extruder.
Melt is supplied to the system at the feed end at some pressure
Po' The flow rate Q through the die at the output end of the
extruder requires a pressure Pe The pressure varies linearly along
the screw.
What can we say about how the output of the system depends on
melt viscosity, screw geometry and screw speed? All of these
variables appear explicitly in Equation 14-6. The only term in the
equation not specified is the pressure gradient aplal. From the
linearity of the pressure we can write this as
(14-9)

ROLE OF RHEOLOGY IN EXTRUSION

455

Equations 14-6 and 14-9 then give us the output Q if the


pressures Po and Pe are known. Because the melt has been assumed to be a Newtonian fluid, Pe is directly proportional to the
flow rate Q and to the viscosity TJ. Therefore,
(14-10)
K is a constant that characterizes the resistance of the die to flow.
For example, if the die were a long pipe of radius R and length L,
from Equation 8-14 K would be 7rR 4 /8L.
The feed pressure Po is determined by conditions in the zone
preceding our system that we have not yet discussed. For the time
being we will treat the situation in which Po is equal to zero, such
as occurs when a melt conveying zone follows a vent zone. For this
special case we can write Equation 14-6 in the form
(14-11)
where we have lumped the product of a number of factors into the
constant C,

7rDbH3 sin2 8
C=----12TJL

(14-12)

Equations 14-10 and 14-11 are easily solved algebraically or, as


shown in Figure 14-4, geometrically. The ordinate in this Figure is
Q and the abscissa is Pe The die characteristic curve, from Equation 14-10, is represented by a straight line through the origin with
slope K/TJ. The screw characteristic curve, Equation 14-11, is also a
straight line, with slope - C. Its intercept on the Q-axis when Pe is
zero is Qd' the drag flow rate. This flow rate is achieved only when
there is no die offering any resistance to flow. The intercept Pmax
on the pressure axis, when Q is zero, is Qd/C. This is the maximum
pressure that can be generated by the extruder.
The intersection of the two characteristic lines is the only point
on the figure where both equations hold and is therefore the
solution of Equations 14-10 and 14-11. It is thus the "operating
point" of the extruder. A figure such as Figure 14-4 is often called
an "operating diagram." It is important to note that the solution is

456

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

SCREW
CHARACTERISTIC

O(DIE) = kp"l1J
O(SCREW) = Od - Cp"

Pmax = OdiC

Figure 14-4. Operating diagram for isothermal extrusion of a Newtonian melt.

a point. For a given combination of screw geometry, die geometry,


and screw speed, the extruder/die combination will only operate at
a single combination of pressure and flow rate. Further, the flow
rate is independent of the melt viscosity. The only way to vary the
throughput in this highly simplified analysis is to change the screw
speed.
We will return to the analysis of extruder operation by use of the
operating diagram below. At this point, however, a few words on
the principle of the analysis are appropriate. What we have done
conceptually is to write the equations for the flow rate-pressure
relationship for two zones: the melt conveying zone and the die. We
then imposed two conditions that must hold in order for these two
zones to be coupled into a single system:
1. The pressure on the melt at the end of the conveying zone is
also the pressure at the die entrance,
2. The flow rate through the extruder is also the flow rate
through the die.

ROLE OF RHEOLOGY IN EXTRUSION

457

Stated this way these conditions seem self-evident; they represent


the principles of conservation of momentum and of mass. These
principles, and also that of conservation of energy, must of course
be satisfied for any model intended to represent a real physical
system. Additional requirements may have to be met, such as
conditions on the velocity distribution rather than the macroscopic
flow rate.
The coupling of the two components-the extruder and the die
-means that it is not possible to specify the operating conditions of
the components independently. The flow rate in the extruder depends on the resistance to flow of the die, for example. In this
simple example there is a unique solution, and it is easy to obtain,
either graphically or algebraically. When one considers the case of a
real extruder, which has numerous zones, the situation is analogous
but mathematically much more complicated. Explicit algebraic solutions are not available, and resort must be made to numerical
solutions. Even these are not necessarily straightforward. Generally
the procedure is to obtain an estimate for each zone and to
compare the coupling conditions, and then to search iteratively for
a solution that satisfies all of the conditions simultaneously. Furthermore, there is no guarantee that the solution found is unique.
There can be multiple solutions, corresponding to different combinations of flow rate, pressure, and temperature that satisfy the
governing equations for each zone and the boundary conditions. In
our simple example this could not occur, because the two straight
lines representing the possible operating situations of each component can intersect at only one point. When one considers the effects
of non-Newtonian viscosity and non-isothermal flow, etc., these
straight lines become curves, and these can intersect at more than
one point. All of the intersections correspond to possible operating
conditions. Such a situation may reflect an instability such as
"surging," in which the extruder output oscillates between different
values.
The operating diagram is a useful tool for analyzing the effects of
extruder and operating variables on extruder performance. Figure
14-5 [2, p. 109] shows the operating lines for two melt conveying
zones. Zone 1 has deeper flights (larger H) and correspondingly
larger drag flow than Zone 2. However, the slope C of Equation
14-12 varies with H 3 , and the characteristic line of Zone 1 is thus
steeper than that of zone 2.

458

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 14-5. Example of analysis by operating diagram. Melt conveying zone 1 has deeper
flights and greater Qd than shallow flighted zone 2. Die A offers larger resistance to flow
than die B. With die A, zone 2 produces larger throughput than zone 1. (A z is at a larger
flow rate than A 1.) The converse is true for die B. Adapted from Ref. 2. Copyright 1985 by
Elsevier Applied Science Publishers Ltd. Reprinted by permission.

Now let us see how these zones behave when coupled with two
different dies. Die A in this figure is represented by a line with a
very small slope, i.e., a small value of K in Equation 14-10. This die
offers severe resistance to flow, and a very large pressure drop is
required to achieve a high flow rate. The operating points for this
die with the two zones considered are denoted as A1 and A2. Point
A1 is at a lower output than A2, even though the drag flow
capability of Zone 1 is much higher than that of Zone 2.
Conversely, if we couple these two zones with Die B, whose
operating line has a very steep slope because it offers little resistance to flow, the output is much higher (Point B1 versus B2) for
Zone 1.
In this example we have looked at the effect of flight depth on
extruder performance. Obviously the effects of any of the other

ROLE OF RHEOLOGY IN EXTRUSION

459

variables that enter equations 14-10 and 14-11 can be analyzed


similarly. More complicated situations can also be examined by this
graphical technique. Stevens [2] gives examples of vented screws, of
combinations with valves and gear pumps, and of screws whose
flight depths are variable. The analysis is valuable for giving quick
insight into the qualitative effects of design and operating variables,
not only into steady state performance but also its sensitivity to
perturbations.
14.2.3 Effects of Simplifying Approximations

In the preceding analysis a number of simplifying approximations


were made. Some of these were geometric, such as neglect of
channel curvature. Another was neglect of "leakage" flow of the
melt through the clearance between the screw flights and the
barrel. Also, the shear heating of the melt was not considered. And
most drastically, the melt viscosity was assumed to be independent
of temperature and shear rate. In this section estimates are given
for the magnitudes of the effects of these approximations.
14.2.3.1 Geometric Factors

The analysis has assumed the simplified geometry of Section 14.2.1


in which the screw channel was described as a wide, shallow strip of
rectangular cross-section. For the most part the lifting of the
geometric approximations has been studied for the case of isothermal flow of a Newtonian fluid. A few studies have extended to
power-law fluids. The results are usually expressed as correction
factors by which the pressure and drag flow terms of the simple
analysis must be multiplied.
If the condition that the channel is wide, i.e., that H jW 1, is
not met, the two terms of Equation 14-6 must be multiplied by
correction factors Fd and Fp to correct for the effect of the finite
channel width. For most applications the corrections are not very
large. The factors Fp and Fd, although not identical, can be approximated by
Fp = Fd = 1 - 0.6HjW

(14-13)

460

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

when H /W is less than 0.6 [1, p. 285]. For the usual range of
extruder screws these corrections are not greater than about 5%.
Another correction that should in principle be applied originates
from the fact that the "tape" corresponding to the screw channel
has oblique, rather than square, ends. The magnitude of the correction factor depends on the helix angle () and on the screw aspect
ratio L/Db' approaching unity as this ratio becomes large. For a
square screw () = 17.6) with an aspect ratio of 8 the correction
factor for pressure flow is 10% [3, p. 220].
Finally, because we have assumed that the channel is shallow,
i.e., that H/Db 1, we have been able to approximate the channel as a straight ribbon. If this approximation is not valid it becomes
necessary to consider the flow in a curved channel. This is a difficult
problem, because it is quite likely that in deep channels, where
H/Db becomes large, H/W also becomes large. It is necessary then
to make corrections for the finite channel width at the same time as
the channel curvature. Tadmor and Klein [3, p. 308] present numerical solutions for the flow of a Newtonian melt for this case. The
curvature correction factors differ from unity by not more than 10%
for channel depth ratios of 0.1 or less, the usual range of melt
conveying sections in plasticating extruders.
14.2.3.2 Leakage Flow

Leakage flow is that occurring in the narrow clearance between the


screw flight and the barrel. As in the case of the flow in the main
screw channel, there are pressure and drag flow contributions to
the total flow. However, because the leakage flow is coupled to the
main flow its analysis is more complicated. A number of approximate solutions have been derived, resulting in a variety of complex
expressions that differ because somewhat different assumptions or
approximations are made in their derivation.
All of the analyses are in agreement that under "normal" circumstances the leakage flow has a negligible effect on overall extruder
throughput. The flight clearance d I is generally about 0.001 times
the barrel diameter Db [1, p. 68], whereas the ratio of channel
depth H to Db is usually on the order of 0.1. The correction for the
effect of the clearance on the drag flow is approximately (1 - d I/ H),
which is equal to 0.99 under these "normal" conditions. The

ROLE OF RHEOLOGY IN EXTRUSION

461

pressure flow contribution is also negligible except at very high


pressures and flow rates. As a rule-of-thumb, clearance effects on
throughput become appreciable only when the clearance is four
times greater than the normal value of 0.001 Db.
However, the additional power required due to leakage flow
cannot be neglected. For the drag flow component it can be
calculated by methods similar to those used to derive Equation
14-7. The result is that the contribution of the clearance flow to the
power, E" is given by
(14-14)
where e is the flight width.
The power calculated in the absence of leakage flow E is given by
Equation 14-7:

We can compare these expressions to estimate the relative contribution of the leakage flow power to the main channel power. The
result is approximately:
(14-15)
Typically elW is on the order of 0.1 [1, p. 384], but Hid, is
perhaps 100, so that their product can far exceed unity. Equation
14-15 overestimates the power dissipated in the clearance, because
it neglects the shear thinning nature of actual melts. A simple way
to correct this is to multiply the right hand side of Equation 14-15
by a factor 1],11], where 1], is the viscosity of the melt at the shear
rate in the flight clearance, and 1] is the viscosity in the screw
channel. Alternatively, Rauwendaal has assumed a power law fluid
and calculated the effect of the power law index on the ratio of the
flight clearance power to the total power for a number of screw
geometries [1, p. 385]. As would be expected, the ratio decreases

462

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

dramatically as the power law index decreases from its value of one
for a Newtonian fluid.
An estimate of the power consumption in the flight clearance is
obviously useful, for example, for determining the size of the motor
required to run an extruder. It must be emphasized once more that
a large fraction of the power consumption goes to raise the melt
temperature and that excessive heating can cause polymer degradation. An estimate should be made by the methods discussed below
of the adiabatic temperature rise in the flight clearance. The
estimate should be made separately for the clearance and the screw
channel, because both the heat generation and the heat conduction
terms are different in the two geometries. The adiabatic temperature increase is a volume average, and the maximum increase
locally at some position in the melt will be higher. If the approximate calculations indicate that the temperature may reach the
degradation range, some experimentation and/or computation by
more sophisticated models is indicated.

14.2.3.3 Non-Newtonian Viscosity

The assumption of a constant viscosity, at 3hear rates corresponding


to those found in an extruder, is not valid even for polymers with
relatively little shear rate dependence. For example, a low molecular weight acetal copolymer, with a melt flow index of 28 dg/min
and a zero shear viscosity of about 5000 poise, shows appreciable
shear rate dependence, with the viscosity dropping to about 3000
poise at a shear rate of 100 s -1. An extrusion grade of this polymer,
with an MFI of 2.8 dg/min, shows much more severe dependence;
its viscosity decreases to about one-third the zero shear value at
100 s -1. These polymers have quite narrow molecular weight distributions. Many other polymers have broader MWDs and correspondingly greater shear rate dependence of viscosity.
The range of shear rates that the melt experiences in an extrusion operation can be estimated to a first approximation by using
the concentric cylinder viscometer equation. Typically the channel
depth H is on the order of one tenth of the barrel radius R. The
angular velocity w in rad/s is 27TS /60, where S is the screw speed

ROLE OF RHEOLOGY IN EXTRUSION

463

in RPM. The shear rate is then

y = wR/H = (27TS/60)R/(R/1O)

=::

For an extruder operating at a typical screw speed of 30 to 60


RPM, the shear rate in the channel due to the rotation of the screw
is therefore on the order of 30 to 60 s - 1. This is well within the
range in which polymer melts show significant shear thinning.
A similar calculation can be made for the shear rate in the gap
between the screw flights and the barrel. Since this gap can be one
fiftieth of the channel depth, the shear rate imposed on the melt in
the gap will be correspondingly higher. The calculation of the
leakage flow through the gap must, therefore, use a viscosity evaluated at the appropriate shear rate, which is on the order of
2000 s - 1. The shear rates imposed on the melt film in the melting
zone are also of this order.
It is not possible to make any general statement about the shear
rate in die flow because of the wide range of possible geometries.
The shear rate can be very low during the extrusion of thick rods or
pipes, but in the melt spinning of fibers the shear rate can be 10 4 or
even 10 5 s-1.
It is clear that in the shear rate range of typical extrusion
applications the viscosity of the melt will vary with shear rate. The
amount of variation depends upon the molecular weight and the
MWD of the polymer. Condensation polymers generally have fairly
narrow MWDs and therefore show smaller variations than addition
polymers. It is important to recognize that this variation is not a
small perturbation. Over the range of shear rates we have discussed, the melt viscosity can vary by several powers of ten. For
quantitative prediction of extruder performance it is essential,
therefore, to take non-Newtonian flow behavior into account.
The characteristic or operating curve for the screw depends upon
the power law index of the fluid and upon the screw helix angle. A
typical result, for a helix angle of 17.66 (lead equal to diameter) is
shown in Figure 14-6. In this figure the flow rate is made dimensionless by dividing by the drag flow rate, Qd. Similarly, the pressure, P, is made dimensionless by dividing by Pmax' the maximum
pressure that the screw can generate. The abscissa, then, is P / Pmax'

464

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1.0~--------------------------------------~
0.9
0.8
~

::l
~

0.7

(!)

::l

~ 0.6

J:

~ 0.5
w

....I

Q 0.4

CI)

w
::E 0.3

1S

0.2

0.1
0~

__4 -__4 -__4 -__

0.1

0.2

0.3

0.4

__

~~4-~4-

0.5

0.6

0.7

__

0.8

__

4-~

0.9

1.0

DIMENSIONLESS PRESSURE

Figure 14-6. Screw characteristic curves for power-law fluids with various values of n.
Adapted from Ref. 1. Copyright 1986 by Hanser Publications, Munich, Vienna, NY;
Distributed in USA and Canada by Oxford University Press. Reprinted by permission.

which in turn can readily be shown to be equal to <1>, the ratio of


the pressure flow to the drag flow.
There is one more point to clarify before discussing the curves of
Figure 14-6. For a Newtonian fluid Pmax is given by

which is obtained by setting Q = 0 in Equation 14-11. What value


of the viscosity should be used for the non-Newtonian case? By
convention, for non-Newtonian fluids the normalization is done
with the viscosity of the fluid at a shear rate equal to 'TTNDb/H, the
linear velocity of the barrel divided by the channel depth.

ROLE OF RHEOLOGY IN EXTRUSION

465

Now let us look at Figure 14-6. There is an obvious effect of the


power-law exponent n on the shape of the screw operating curve.
At any given pressure the throughput decreases as n decreases, i.e.,
as the fluid becomes more strongly shear-thinning. Conversely, the
pressure generated at a given throughput is greatly reduced compared to the Newtonian case. The efficiency of the screw as a melt
pump is also decreased by the shear rate dependent viscosity. Booy
[7] has calculated the efficiency for power-law fluids. As an example, the maximum efficiency for a power-law fluid with an exponent
n = 0.5 is about 22%, compared to 28% for a Newtonian fluid.
Furthermore, the throughput for the maximum efficiency also varies
with n, decreasing from 65% of the drag flow rate at n = 1 to 40%
at n = 0.3.
It is interesting to note that even the unrestricted drag flow, when
the pressure is zero, is smaller for a power-law fluid than for a
Newtonian fluid. At first glance this is surprising, because the
one-dimensional drag flow result is independent of the fluid viscosity or of its shear rate dependence. The reason for the effect of the
fluid rheology in this case is that the flow in the extruder screw is
two-dimensional. Thus, shearing associated with the transverse flow
alters the viscosity throughout the channel, and this in turn affects
the drag flow rate.
For quick estimates of the operating curve it is desirable to have
approximations that are easy to use. The simplest estimate, perhaps, is that of a Booy [7]. He calculates a reference shear rate Yo,
which is equal to the Couette flow shear rate in the channel
between the barrel and the screw.
(14-16)

The reference viscosity 'Tlo is the viscosity of the power-law fluid at


the shear rate Yo. (This is the same procedure, described above,
that was used to normalize the operating curves.) Booy then finds
that the operating curve for a power-law fluid can be approximated
by a straight line whose intercept on the flow rate axis at zero
pressure is the drag flow. The slope is steeper than for the reference case of a Newtonian fluid with viscosity 'Tlo; the ratio of slopes
is approximately lin over wide ranges of n and operating condi-

466

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


1.0~--------------------------------------~

I:::l

0.8

Q.

J:

(!)

:::l

~ 0.6
J:
I-

en
en
w

2!
enz

0.4

::!:

0.2

O~

____

~~

____

0.1

______

0.2

0.3

____

~~

____

0.4

__

0.5

..

~~

0.6

DIMENSIONLESS PRESSURE

Figure 14-7. Example of approximations proposed for screw characteristic for power-law
fluids, for a fluid with n = 0.5. Solid line is the exact characteristic curve from numerical
computation. Dashed line is Booy's approximation [7], with slope equal to -lin. Dotted line
is Rauwendaal's approximation [1], which is Equation 14-17.

tions. Booy's approximation is illustrated for the case n = 0.5 in


Figure 14-7.
A slightly more complex approximation for the characteristic
curve of a power law fluid worked out by Rauwendaal [1, p. 301] is
given in dimensionless form by Equation 14-17. An example of
Rauwendaal's approximation is also shown graphically in Fig. 14-7.
Q

f == Qd

4+n
( 3 )
-5- - 1 + 2n <I>

(14-17)

In order to determine the operating point of the extruder we


need also the characteristic curve of the die. Die flow of nonNewtonian fluids has been discussed in Chapter 8. For shear-thinning fluids, such as most polymer melts, the characteristic is a curve
that is concave upward, with the flow rate increasing faster than

ROLE OF RHEOLOGY IN EXTRUSION

467

linearly with increasing pressure. The combination of the curvatures


of the die and extruder characteristics shifts the operating point to
a lower pressure than for a Newtonian melt. The effect upon the
throughput is not so easily predicted; it depends on the details of
the shapes of the die and extruder characteristics, because the
effects of their curvature tend to oppose each other.
14.2.3.4 Non-isothermal Flow

We now consider another of the simplifying approximations made


so far, namely, that the flow in the extruder channel is isothermal.
This cannot be true in practice. It was pointed out in the discussion
of the efficiency of the screw extruder as a pump (Equation 14-8)
that only a small fraction of the total power is actually used to
produce pressure to pump the melt against the resistance of the
die. The major portion of the work done is dissipated as heat. Some
of this heat can be removed by conduction to the barrel and to the
screw, both of which can be cooled by circulation of a fluid. The
remainder, however, raises the temperature of the melt and decreases its viscosity.
The magnitude of the temperature rise and the associated change
in the screw characteristic curve depend on the amount of heat
generated by viscous dissipation. But they depend also on the
amount of heat lost by conduction. It is possible to write
the equations describing these processes, also taking into account
the non-Newtonian behavior of the melt. The equations are complex and can only be solved numerically. The numerical solutions
depend on the nature of the boundary conditions assumed; for
example, one can assume that at the barrel wall the melt has some
specified temperature.
Solutions of these equations are given in the literature [1,3,8].
For sufficiently simple boundary conditions and with some other
assumptions the reSUlting screw characteristic curve depends on
three parameters: the screw helix angle, the exponent of the assumed power-law fluid behavior, and a modified Brinkman number
{3, defined by:
(14-18)

468

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Here b, K b , and n are parameters governing the rheological


behavior,

The parameter k is the thermal conductivity of the melt. The other


parameters of Equation 14-18, the channel depth H and the
relative velocity of the barrel and screw, v bz ' are parameters of the
extruder design and its operation.
The dimensionless parameter {3 is proportional to the ratio of the
amount of heat produced by shear to the amount removed by
conduction and depends upon the amount of viscosity variation
caused by a given temperature rise.
When {3 becomes greater than zero the screw characteristic curve
departs from its isothermal non-Newtonian shape. At {3 = 1 the
departure becomes marked, as shown in Figure 14-8, and becomes
increasingly so at higher values of {3. The flow rate can increase or
decrease, compared to the isothermal case, depending on the
pressure gradient.
The Brinkman number incorporates factors dependent upon the
extruder geometry (barrel diameter and channel depth) and operating condition (screw speed) as well as parameters that characterize
the melt viscosity. This dependence of the Brinkman number re0.5

1Il"t""------------------,
17.7 0

(J =

I>o.~ ....

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

........ ........f3
f3

........

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

Figure 14-8. Effect of Brinkman number on screw characteristic. Adapted from Ref. 8.
Copyright 1977 by McGraw-Hill Publications. Reprinted by permission.

ROLE OF RHEOLOGY IN EXTRUSION

469

flects the fact that the relative importance of heat generation and
conduction depends on the size or surface-to-volume ratio of the
extruder. In a small laboratory extruder, in which melt thickness is
small, heat conduction is favored, and the extrusion conditions may
be close to isothermal. The same operation conducted on a large
production machine, with large melt thickness and poor heat conduction, will be much more nearly adiabatic, resulting in a considerably larger temperature rise.
As indicated above, the relevant equations are so complex and
their solutions so specific to assumptions and boundary conditions
that if a relatively precise estimate of the extruder operation is
needed, it is probably best to use the numerical methods incorporated into the commercially available extruder models. However,
approximate, simple solutions are useful in some instances. At the
least they can provide an estimate of the need for more exact
computation.
For example, McKelvey [9, pp. 256-265] has solved the case of
the adiabatic extrusion of a Newtonian fluid whose viscosity varies
exponentially with temperature and has presented the solution
graphically. This is useful because many polymers degrade when
exposed to too high a temperature, and even a rough estimate of
the temperature rise during extrusion can suggest the need for
caution in selecting extrusion conditions. Middleman [8, pp. 131137, 144-149] gives illustrative examples of the use of this solution
and extends it to the non-Newtonian situation by a well defined,
albeit tedious, approximate procedure.
The adiabatic temperature rise can be calculated simply from
Equations 14-19 for the case of open discharge extrusion, i.e., pure
drag flow.

dT

27r'T] bND b L ( 1 + 3 sin 2 0 )


pCH 2
cos 0 sin 0
(lib )In(l

+ Z)

(14-19a)
(14-19b)

In this Equation p is the melt density and C its heat capacity; b is


the exponent of the temperature dependence of viscosity as defined
above.

470

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

14.2.4 Solids Conveying and Melting Zones


14.2.4.1 Feeding and Solids Conveying

Until now we have discussed the extruder only from the point of
view of its function as a melt pump. In the large majority of cases,
however, we are interested in the process of plasticating extrusion.
Here the material fed to the extruder is a solid, generally granular,
which must be compacted and melted by the action of the extruder.
The mechanics of these processes is not as well understood as that
of the melt conveying process discussed above. The analysis is
complicated by the diversity of operating conditions required for
different materials and by differences in melting mechanisms from
one material to another.
The flow of granular solids into and within the extruder is
affected by the friction between the particles and the friction of the
particles against the metal surfaces with which they come into
contact. These frictional forces are highly variable. They depend
upon the materials, upon conditions such as temperature and
humidity and the presence of lubricants and on the size and shape
of the granules. Some of these factors are not always controllable or
may be difficult to simulate in laboratory tests. Rauwendaal
[1, p. 170] reviews the literature on the measurement of the coefficient of friction.
In practice there are additional complications to consider, both
from the viewpoint of theoretical analysis and of operation of an
extrusion process. The solid feed to the extruder is rarely uniform
in size, shape, and friction coefficient. The feed may contain a
variable amount of reground solid material from various sources
such as off-specification product, trim from blow molding and film
extrusion processes, and sprues and runners from injection molding. The presence of these irregularly shaped regrinds will affect the
flow of the solid feed. In blending processes the two or more
feedstocks will in general have different forms, densities, and friction coefficients. Not only the flow but even the composition of the
feed may then be variable because of the possibility of segregation
of components during feeding.
Segregation may be a problem even when one is extruding virgin
powder from a polymerization reactor, because the powder may

ROLE OF RHEOLOGY IN EXTRUSION

471

have a range of particle sizes. There may even be systematic


variations of molecular weight with particle size, presumably because of differences in precipitation kinetics with molecular weight
[10, 11]. Segregation during feed can then cause unacceptable fluctuations of the extrusion process and of product quality.
The simplest mechanism for supplying an extruder with feed is by
gravity flow from a hopper. For materials that are moisture sensitive the hopper can be arranged to permit blanketing with dried air;
the hopper can even double as a drying device. The most serious
problem encountered with hoppers is that the interparticle friction
can be so high that the feed can "bridge," that is, pack together, to
form a barrier that interrupts flow and prevents feeding completely
or causes a variation of the pressure at the extruder throat. The
pressure at the beginning of the solids conveying section of the
screw has a strong effect on the performance of the extruder, and
variations in the pressure can cause large changes in extrusion rate,
power consumption, and melt temperature rise. Vibration of the
hopper or gentle agitation of the feed in the hopper can be used to
prevent bridging.
A more positive way to minimize bridging is to equip the hopper
with a screw to convey the feed to the hopper throat. Such a screw
can also act as a force feeder, generating considerably more pressure than gravity can. Examples of the successful use of force
feeding for extrusion of difficult materials, such as scrap film, are
given by Kruder and Kim [12]. They show that output can be
increased by 50 to 100% and surging can be eliminated by force
feeding.
Gravity or force feeding results in a method of operation called
"flood feeding." The throughput of the extruder is then determined
solely by its melting and pumping capacity. An alternate mode of
operation is "starve" or "metered feeding." In this case the
throughput is governed by the rate at which the feed is metered to
the extruder. Gravimetric or volumetric devices can be used for
metering. Twin screw extruders are usually operated in a metered
(starved) mode.
Moderate starve feeding, at about 90% of the floe::! fed maximum
throughput, can have numerous beneficial effects [13]. It reduces
screw torque, specific power consumption, pressure fluctuations at
the die, and melt residence time. The degree of improvement will

472

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

depend upon the particular combination of material and extruder,


but appears to increase with increasing extruder size. There is a
reduction in throughput, but in practice this can be offset by
operation at a higher screw speed.
Once the feed material enters the screw it is conveyed forward by
the frictional forces between the particles and the extruder surfaces. Depending upon the details of the feeding arrangement and
the material properties, the feed can fill the screw almost immediately and be compacted into a solid plug by the generated pressure.
Particularly in the starve feeding mode, the screw can remain
partially filled for some distance before compaction occurs.
Up to this point the behavior has been dependent on too many
unknown or variable factors to be analyzed generally. Once the
feed has been compacted into solid form, the conveying process is
amenable to analysis using a technique of Darnell and Mol [14].
The solid is assumed to flow down the screw channel as a plug. The
velocity is proportional to the screw speed and is determined by the
balance of forces and torques exerted on a section of the solid plug
at the screw and barrel surfaces, and by the pressure gradient down
the plug.
The friction at the barrel wall is responsible for the forward
motion of the solid plug. The friction on the screw is a retarding
force. It is important, therefore, to insure that the friction coefficient at the barrel wall is greater than that on the screw. For this
reason the hopper is often cooled, and the rear section of the barrel
is held at a relatively low temperature, in order to avoid premature
melting of the polymer at the barrel surface, which would result in a
low effective friction force. A positive method to obtain high barrel
friction is to roughen the barrel surface. Also, the barrel can have
grooves machined into it. The possible disadvantage of excessive
barrel friction is that very high pressures can be produced and
excessive power consumption and heat generation may occur.
14.2.4.2 Melting Zone

Heat generated by friction, and conduction of heat from the barrel


wall, combine to raise the temperature of the solid bed as it moves
down the extruder. When the temperature reaches the melting
point of a crystalline polymer or the glass transition temperature of

ROLE OF RHEOLOGY IN EXTRUSION

473

FLIGHT

CIRCULATORY ~~+I
FLOW OF
PREVIOUSLY
MELTED
POLYMER

Figure 14-9. Sketch of the melting zone as envisioned in the Maddock-Tadmor model.
Adapted from Ref. 3.

an amorphous polymer, a melt film forms at the barrel surface. The


shearing of this melt film generates more viscous heat, which is
conducted into the solid bed and melts more polymer.
A qualitative understanding of the details of the melting mechanism is based on the observations of Maddock [15]. He devised a
technique for rapidly freezing the material in an extruder and
extracting it with the screw. The visible difference between the
compacted solid and the solidified melt made it possible to follow
the course of melting. These observations were the basis of the
subsequent Tadmor quantitative theory of the melting zone [16];
numerous variations have been proposed since. These have been
compared and reviewed by Lindt [17], who concluded that" ... the
development of the melting theory for the Maddock mechanism as
an engineering analytical tool is virtually complete."
The idealized mechanism of melting based on Maddock's observations is shown schematically in Fig. 14-9. As melting proceeds the
melt film becomes thicker. Eventually its thickness becomes comparable to the clearance between the screw flight and the barrel. At
this point further melting causes accumulation of a melt pool at the
rear of the screw channel, adjacent to the pushing side of the flight.
With further melting the melt pool width increases and that of the
solid bed decreases correspondingly, till all of the polymer has
melted.

474

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The effect of the rheological behavior of the molten polymer on


the melting process is then simply due to the proportionality of the
viscous dissipation in drag flow to the viscosity. We have encountered this previously in the discussion of the melt conveying zone
(Equation 14-7). Because the viscosity depends on temperature and
shear rate a quantitative statement requires numerical analysis of
the coupled viscous flow and heat transfer processes. The geometry
of the extruder also has an important effect on the details of the
melting kinetics. This is because the ratio of heat conduction to
heat generation varies with extruder size. Also, the flight clearance,
which determines the shear rate in the melt film, also generally
varies with the screw diameter. Extrapolation or scaleup of the
melting zone is, therefore, particularly difficult.
Observations have been made of phenomena not included in the
idealized Maddock mechanism, and elaborations of the original
Tadmor model have been proposed to account for these phenomena or other perceived deficiencies of the model. In some instances
polyviny1chloride has been observed to form a melt pool at the
trailing rather than the leading flight. It has also been noted that a
melt film can form at the screw surfaces as well as at the barrel
surface. The cross-section of the melt film can be tapered, not
rectangular, increasing in thickness towards the melt pool. There
may not even be a melt pool, but rather a solid bed surrounded by
melt on all sides. The details of the motions can also differ; the bed
can be rigid or deformable, and many types of flow pattern in the
melt film and in the melt pool are possible.
The modelling of the melting process has been reviewed by Lindt
[17]. He concludes that all of the models give a reasonable estimate
of the solid bed width as a function of distance along the extruder.
This is a fortunate circumstance for extruder simulation, but it
means that accurate prediction of the profile is not a critical test for
the validity of a model. The pressure development in the melting
zone is much more sensitive to the details of the model. According
to Lindt the deformable bed models seriously underestimate the
pressure, while rigid bed models do much better. The best estimates of both melting and pressure profiles is achieved by combining a rigid solid bed with circulating cross-channel flow in the melt.
This is important because of the interaction of the pressures in the
various zones of the extruder.

ROLE OF RHEOLOGY IN EXTRUSION

475

Further improvements of these models, such as would be achieved


by incorporating melt viscoelasticity, would complicate the computation more than would be justified by possible improvement of the
simulation.
One of the assumptions of the Maddock model that may not
always be valid deserves separate discussion. Figure 14-9 shows the
solid bed continuously diminishing in width but always retaining its
identity as a coherent solid. In fact the forces exerted on the bed
during its passage can cause it to break up into fragments of various
sizes. If the fragments are small and are suspended in the melt
pool, their rate of melting will be considerably lower than when
present as a solid bed. Breakup will therefore affect the temperature of the melt delivered to the conveying zone. It can also induce
instability of the operation, because unmelted fragments can accumulate and choke off flow in the transition between the melting and
conveying zone. In any case the pressure development in the
melting zone, and therefore also downstream, will be affected by
bed breakup.
There does not appear to be any theory for the prediction of the
occurrence of bed breakup or for its effects. Experimentally, it
appears to depend on the nature of the polymer, the pellet size,
and the screw geometry [3, p. 88].
Recognition of the advantage of preventing bed breakup has led
to the invention of a wide array of screw designs whose purpose is
to separate the melt pool and the solid bed. By confining the bed,
the probability of breakup is minimized. The performance characteristics and the manufacturing problems associated with a number
of these designs are discussed by Rauwendaal [1, pp. 402-415], who
also describes some designs in detail and analyzes their melting
behavior. Analyses are also to be found in other publications
[18-20]. In general it can be said that proper design of a barrier
screw can result in improved melting capacity and stability of
extrusion. A price that one pays, in addition to the increased cost
of manufacturing a more complicated design, is loss of flexibility of
operation. In other words, the screw will give improved performance under the design conditions, compared to the simple screw.
Its performance may be more sensitive, however, to variations from
the design conditions. The advisability of a specialized design
depends, then, on whether the extruder is intended to operate

476

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

exclusively for a specific purpose or whether it is required for a


range of applications.
14.2.5 Scale-up and Simulation

We have described in the preceding sections the elements of


models for the essential zones of a plasticating extruder. Many of
these models do creditable jobs of simulating or analyzing the
performance of the zones that they treat. Further, these individual
zone models have been combined in order to simulate the overall
performance of an extrusion operation, including throughput, profiles of pressure and temperature along the extruder, and even
some indication of stability of operation and uniformity of the
extrudate. This section addresses the question of how one can use
the analyses and models for practical applications.
14.2.5.1 Scale-up

Very often the problem one faces is not that of a complete screw
design but of scaling up a satisfactory extruder, or perhaps of
modifying an existing design that is partially satisfactory but has
some defect that is well diagnosed. In these cases it may not be
necessary to use the full power and complexity of the simulation
models. The analytic relations for the various zones given in the
preceding sections (and in the references) can be used directly.
There is no unique set of scaling relations applicable to all
situations. The targets to be met can differ widely. If one is
extruding a shape whose dimensions are critical, constancy of
throughput is obviously extremely important, and the scale-up strategy must not jeopardize it. If, on the other hand, the operation is
one of pelletizing, then this is not so important a consideration, but
maximizing throughput may be. If the polymer is very sensitive to
thermal degradation, the scale-up must not increase melt temperature and/or the residence time. The compromises that must be
made in scaling up arise because the different process elements that
occur in extrusion depend differently upon the dimensions. For
example, the drag flow rate in the metering zone is proportional to
the channel depth H, but the pressure flow rate varies with H3.
When H is increased for scale-up, the two flow rates will increase

ROLE OF RHEOLOGY IN EXTRUSION

477

by different amounts unless some other dimension is changed as


well. But that change will in turn affect some other process variable.
For this reason a wide variety of scale-up relations have been
proposed, starting with that of Carley and McKelvey [21]. Stevens
[2] tabulates analyses of ten possibilities and discusses their relative
merits and disadvantages. Rauwendaal [22] reviews seven older
methods and proposes two new ones for different situations.
The most commonly used scale-up procedure [22, 23] is attributed
to Maddock [24]. The extruder L/D ratio is kept constant, and the
channel depth is increased by the factor d l/2 when the diameter is
increased by a factor of d. The screw speed N is decreased by d l/2
upon scale-up. The result is that the shear rate experienced by the
melt is unchanged. The throughput (more precisely, the drag flow
rate) increases as the square of the diameter ratio. According to
Chung [23] this scale-up strategy is deficient because the melting
rate only increases with the 1.75-power of d, compared with d 2 for
the pumping rate. Rauwendaal [22] counters, however, that this is a
problem only when the melting mechanism is predominantly heat
conduction from the barrel rather than viscous dissipation. This
controversy illustrates that the analysis of extruder performance is
not a routine procedure but requires understanding of what is
occurring.
Another shortcoming of the Maddock scale-up is that the specific
energy consumption (SEC) increases. This causes increased viscous
dissipation and a melt temperature rise. Chung [23] suggests that
both problems, inadequate melting capacity and excessive energy
dissipation, can be alleviated by shortening the length of the metering zone and increasing the length of the melting zone. Fine tuning
of such an adjustment might be a good problem to consider using
the complete simulation models.
The other scale-up strategies in the literature also have various
advantages and disadvantages. For a detailed discussion the reader
is referred to the original article [22].
14.2.5.2 Simulation

In principle one could use a simulation program to make an


absolute prediction of how a new polymer would be extruded by a
screw of arbitrary design at some prescribed conditions such as

478

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

screw speed and barrel temperature profile. In reality this would be


economically a very risky procedure for at least two reasons. The
first reason is that it is unlikely that the simulation would be
sufficiently accurate for commercial application. Secondly, even if
by some chance the simulation were completely accurate for the
chosen conditions and screw design, it is even less likely that these
would have been selected as the optimum, or even acceptable, ones.
The reasons for inaccuracy of simulation can be summarized as
follows:
1. Inadequacy of the models, especially for simulation of solids

conveying or of bed breakup in the melting zone.


2. Inadequate knowledge of parameters such as the friction
coefficients, the bulk compressibility of granular solids, and
the transient non-isothermal rheological behavior.
3. Excessive sensitivity of simulation to small variations of
boundary conditions, for example due to uncertainty in the
precise location of a thermocouple in the barrel wall.
It must also be remembered that the commercially available
simulation algorithms are presented to the user as "black boxes,"
requiring certain input data and returning results. Changes can be
made in the algorithm by the software vendor without the knowledge of the user, and in some instances these will change the results
of the simulation. Such changes during the course of a series of
simulations, as discussed below, can produce undesirable consequences.
Despite these limitations the simulations can be very valuable in
the expensive and time consuming process of designing a new screw
by trial-and-error and in optimizing the process conditions for a
new material in an existing extruder. A realistic simulation requires
accurate material property data. Rheological properties, melt density and compressibility, and thermal data for both the solid and the
melt can all be measured with sufficient accuracy. The bulk and
friction characteristics of the solid are most difficult to measure,
especially the conditions for solid bed breakup. Probably the safest
strategy is to measure the material properties that can be measured
accurately and to obtain extrusion data over a range of conditions.
Successful simulation of these extrusion experiments should then

ROLE OF RHEOLOGY IN EXTRUSION

479

lead to a consistent set of the material properties that are difficult


to measure. These properties can then be used with greater confidence for the prediction of the performance in the screw to be
designed.
Similarly, if performance over a range of known conditions can
be simulated successfully, it is likely that the effect of a change of
conditions is predictable. At first blush this may not seem such a
complex task as to require a sophisticated computer program. This
is probably true if one is only interested in overall performance
measures such as throughput. If, on the other hand, it is important,
for instance, to limit the time of exposure of the melt to temperatures above some critical value, simulation is probably the best way
to design a screw and/or select extrusion conditions to meet such a
requirement.
The second difficulty alluded to refers to the fact that a single
simulation predicts performance only at the precise conditions
stated. But many independent variables can affect extrusion, and in
general it is necessary to know something about the effect of each
one. Clearly one cannot use simulation blindly and hope to arrive at
a unique optimum result. Rather, one must think of a simulation as
an experiment and use all of the tools of an experimentalist in
devising simulation conditions and in analyzing the results.
Tadmor and Klein [3, Sect. 11.2, pp. 454-463] give an example
of the analysis of a problem of minimizing extruder discharge temperature for a range of six operating variables-screw speed,
channel depth, metering zone length, throughput, and barrel and
input temperature. A factorial experiment requiring 228 simulations
was devised, and the results were analyzed by statistical multiple
regression analyses. They also give references to other analyses of
simulations and of experiments. Anyone thinking seriously of using
simulations is advised to read these analyses, and perhaps to retain
the services of a statistician as well.
It should be pointed out that the problem of screw design will be
even more complex if it involves many more parameters than arise
in the example cited. The experimental design may also be subject
to some constraints, so that not all of the parameters can be varied
independently. For example, in designing a screw for an existing
extruder barrel, the diameter and length are fixed. One cannot vary
the length of the metering zone, say, without affecting that of the

480

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

other zones. Or, the existing extruder may have a drive with a
certain power or torque limitation. The simulations should certainly
be run according to a design and analyzed statistically, as Tadmor
and Klein show. The results of the regression analyses can then be
subjected profitably to optimization techniques such as "linear
programming" [25] that permit imposition of constraints.
One last word about the use of simulation is in order. The results
should be tested for sensitivity to small variations of the simulation
parameters, especially those that are difficult to measure or likely to
change with time. One should be suspicious of a simulation result in
which the more important responses, e.g., throughput or melt
temperature, are very sensitive to small variations of the friction
coefficients or the location of a thermocouple. Such sensitivity
might well imply unstable operation. In such a case it is advisable to
modify the design to sacrifice throughput, for instance, to gain
insensitivity to operating parameters. As another example, it would
not be sensible to design a screw whose performance varied markedly with a small increase of flight clearance, because the inevitable
abrasion during running would cause rapid deterioration of the
performance of such a screw.

14.3 MIXING, DEVOLATILIZATION AND TWIN SCREW EXTRUDERS

14.3.1 Mixing

Mixing is one of the important functions of an extruder. Even if


only one component is being extruded, melt subjected to different
conditions of viscous dissipation and heat transfer, and therefore
having different temperatures, must be homogenized in all but the
most crudely controlled processes. However, as the mixing literature makes clear, it is a difficult operation to quantify and about
which to generalize. The measure of "goodness" of mixing and the
process steps required to achieve the desired level depend on what
is being mixed and for what purpose.
For example, if we wish to disperse carbon black in a polymer for
the purpose of coloring it, it is only necessary to distribute the
particles with sufficient uniformity to give adequate appearance.
However, if our purpose is to make an electrically conducting

ROLE OF RHEOLOGY IN EXTRUSION

18

481

MINOR COMPONENT

MAJOR COMPONENT

DISTRIBUTIVE
MIXING

DISPERSIVE AND
DISTRIBUTIVE
MIXING

Figure 14-10. Schematic comparison of distributive and dispersive mixing.

mixture, the carbon black aggregates must be broken down to a


sufficiently small size so that they can form a continuous network
throughout the material (Fig. 14-10). A similar example is given by
McKelvey [9, p. 303] for the case in which the carbon acts as an
ultraviolet light absorber to protect the polymer from photodegradation.
In discussing mixing it is useful to distinguish between two mixing
processes: "distributive" and "dispersive." The aim of distributive
mixing is to homogenize the composition so that any volume element drawn from the mixture has a composition as close as possible
to the average. The allowable spread of compositions and the
minimum size of the volume element depend upon the intended
application of the mixture. In dispersive mixing one desires also to

482

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

minimize the size of particles of the minor component. The term is


most commonly applied to mixtures of solid particles in a liquid. In
the above example, the efficiency of dispersive mixing increases
with increasing melt viscosity.
However, excessive stress can be harmful, for example when it
leads to the attrition of glass or other rigid reinforcing fibers during
compounding. This results in partial loss of the reinforcing action,
as well as in abrasive wear of the extruder. One possible mechanism
for attrition is the bending of fibers that are anchored at one end in
the solid bed and exposed to the stress of the melt film in the
melting zone [26]. This mechanism is avoided by the common
practice of feeding chopped fiber through a side port past the
melting zone.
Dispersive mixing is applicable not only to mixtures containing a
solid component. As discussed in Chapter 11, over some concentration range, mixtures of immiscible polymers can form discrete
dispersions. The disperse particle size depends on the stress on the
matrix polymer (through the Weber Number) and on the ratio of

n=O

n=2

n=3

1000

Figure 14-11. Laminar shear mixing between concentric cylinders for various initial orientations of minor component, as a function of total strain. Adapted from Ref. 9. Copyright
1962 by John Wiley & Sons, Inc. Reprinted by permission.

ROLE OF RHEOLOGY IN EXTRUSION

483

the viscosities of the components. In this case the smallest disperse


size is not achieved by increasing the matrix melt viscosity but by
matching the viscosities of the two phases.
The primitive mechanism for distributive mixing of polymer melts
is laminar shear during processing. Ottino and Chella [27] have
reviewed the theory of laminar mixing and give numerous literature
references. Figure 14-11 shows how a shearing deformation stretches
out a fluid volume element and increases the area of contact with
the surrounding fluid. If the shear is applied to an array of layers of
fluid, it will reduce the thickness of the laminae and the distances
between them.
However, in practice laminar mixing alone is not adequate. First,
its effectiveness depends on the orientation of the fluid element
with respect to the shear direction, as shown in Figure 14-11. Also,
the fine structure of the mixture may not be suitable for all
applications. It may be acceptable for pigmentation, for example,
but useless for the elastomer toughening of a brittle thermoplastic.
Furthermore, the process is relatively inefficient. For an initially
random array of interface orientations, the increase of interfacial
area in such a process is directly proportional to the shear strain y
[28].

A/Ao

f(y)

y/2

(14-20)

A way to increase the efficiency of mixing is to reorient the


direction of shear. Quantitatively, if a mixer were divided into N
sections, each of which imparted a shear strain y / N and between
each of which the fluid elements were reoriented randomly, the
interfacial area ratio would then be [28]

A/Ao

[y /(2N)t

(14-21)

Consider the case of y /2 equal to 12; for N = 2 the area ratio


increases to 36, for N = 3, to 64, etc. In addition, of course, the
reorientation eliminates possible planes of inhomogeneity that might
promote weakness and excessive permeability.
The deformation in a single screw extruder is primarily shearing.
It is not, therefore, inherently a good mixing device. A large variety
of screw elements designed to improve mixing, dispersive and

484

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

distributive, have been used. These are described, and the pressure
drop they require calculated, by Rauwendaal [1, Section 8.7]. He
also discusses static or motionless mixers, which are elements added
at the exit of the extruder for distributive mixing [1, Section 7.7].
The mixing effectiveness of several designs has been analyzed
theoretically by Erwin and Mokhtarian [29] in terms of the reorientation concept.
One of the reasons for the introduction of twin screw extruders is
to improve mixing. These are discussed briefly in Section 14.3.3.

14.3.2 Devolatilization

Devolatilization is often a necessary step during extrusion. It may


be required to remove residual solvent or monomer after a polymerization reaction or to advance a reversible condensation reaction during extrusion by removal of volatile reaction products. Some
devolatilization can occur by backward transport of gases to the
hopper from the compaction in the solids conveying zone. In order
to assure removal of large amounts of volatiles, extruders are
equipped with vent zones.
The technique [30] requires that the melt conveying zone be
interrupted by a section of the screw with a deeper flighted vent
zone. The melt pressure drops to zero in the vent zone, because it
can convey a larger amount of melt than is delivered to it and is
therefore only partially filled. The melt forms a pool at the pushing
flight of the screw, leaving a thin film of melt on the barrel, as
shown in Figure 14-12. Becuase the rate of diffusion increases
rapidly as the film thickness decreases, volatiles can diffuse from
the melt into the gas space of the partly filled channel. Some
devolatilization can also occur from the face of the melt pool. The
conveying action of the screw circulates the melt pool and interchanges melt from it to the film phase.
The volatiles that accumulate in the gas space are removed from
the extruder through a vent port in the barrel, usually by pulling a
vacuum. If the concentration of volatiles in the feed is high, the
application of a vacuum can result in foaming or bubble formation.
The breaking of these bubbles is a more efficient mechanism for
devolatilization than diffusion. In fact, an inert volatile material

ROLE OF RHEOLOGY IN EXTRUSION


POOL

485

--BARREL

.::
.

:: ..
: .

: :

: : ..
: : :

. : :

.
. //
/Fi~ 1
/

. /~ANC
../ ~
.
'/ '

/'

.-

/'

: :

/'

pool in ve
e o.t..extruder, showing surfaces for devolatilization by
evaporation. Adapted from Ref. 30. Copyright 1983 by Hanser Publications, Munich,
Vienna, NY; Distributed in USA and Canada by Oxford University Press. Reprinted by
permission.

such as water can be added as a stripping agent to aid in the


removal of an undesirable component.
As with mixing, devolatilization can be accomplished more readily by a properly designed twin screw extruder than by the single
screw device we have discussed. We now turn to a brief description
of twin screw extruders.
14.3.3 Twin Screw Extruders

Twin screw extruders are designed to overcome two disadvantages


inherent to single screw extruders. One is that in single screw
machines the solids conveying and melting processes are coupled to
the melt conveying or pumping zone. The throughput-pressure
relation therefore depends on the behavior at the feed end of the
extruder. As we have discussed, this is relatively poorly controlled,
because it depends on the highly variable friction coefficient, especially when compounding with low melting, low viscosity additives.
The second disadvantage is that the mixing in single screw extruders is relatively poor.
The price paid for the advantages of twin screw extruders is the
increased cost of the more complex design of the components.
There is a wide variety of design types. Figure 14-13 illustrates the
possible combinations of flow patterns achieved by variation of the
direction of rotation, the degree of intermeshing, and the flight

%
w

LENGTHWISE AND
CROSSWISE OPEN

a:

OW
z:E

~(I)

z
%

c:J

-.

<W
~!z

a: a:

~:E

<(I)
_W

LENGTHWISE AND
CROSSWISE OPEN

LENGTHWISE AND
CROSSWISE OPEN

-----

o.

~
- - - -

II

98

9A

12
-

lOA

--

THEORETICALLY
NOT POSSIBLE

~ 8
- .
~
~
~
IOB~

Ii

~
THEORETICALLY
NOT POSSIBLE

THEORETICALLY
NOT POSSIBLE

CO-ROTATING

COUNTER-ROTATING

THEORETICALLY POSSIBLE
BUT PRACTICALLY
NOT REALIZED
5

LENGTHWISE OPEN
AND
CROSSWISE CLOSED 3

LENGTHWISE OPEN
AND
c:J
~! CROSSWISE CLOSED 7

...Il:

-~

~~

...I

~(I)

LENGTHWISE AND
CROSSWISE CLOSED

SYSTEM

Figure 14-13. Types of twin screw arrangement. Adapted from Ref. 31. Copyright 1981 by John Wiley &
Sons, Inc. Reprinted by permission.

i~a:

c:J

a:

:E

z
%
(I)

c:J

SCREW
ENGAGEMENT

G>

C/)
C/)

o()

:Il

."

C/)

()

~:::!

."

Iii

C/)

:Il

=i

~
z>

:Il
:I:

!:i

s::
m

Cl)

it

ROLE OF RHEOLOGY IN EXTRUSION

487

0,
~

TRANSPORT DIRECTION

Figure 14-14. Leakage flow paths in twin screw extruder. Adapted from Ref. 6. Copyright
1978 by Elsevier Science Publishers. Reprinted by permission.

design [31]. Even this classification does not exhaust the possibilities. For non-intermeshing screws the relative placement of the
flights-staggered or matched-also affects performance [32].
Further flexibility of design is afforded by modular barrel and
screw elements. The use of modular barrel elements permits the
optimal placement of vent and addition ports. The screw elements
include kneading discs for mixing, reverse elements for sealing vent
sections, and various numbers of parallel flights for different applications. Some manufacturers even have models that can be converted from counter-rotating to co-rotating operation by moving a
lever.
The flow pattern in a twin screw extruder is, understandably,
considerably more complex than in a single screw machine. For
example, Janssen [6] identifies four different leakage flows, shown
in Figure 14-14, in a closely intermeshing twin screw extruder. One
of these is the flow between the flight and the barrel wall that also
occurs in single screw extruders. Another, called calendar leakage,
occurs between the flight of one screw and the channel of the other.

488

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

There are two flows, one essentially transverse and the other radial,
through the gap between the flanks of adjacent flights.
The complexity of the flow and the large number of possible
designs have limited the possibility of a general analysis of twin
screw extrusion operation comparable to that discussed above for
single screw extrusion. We refer the reader to the specialized
literature on this subject [1,6,31]. It should be noted also that few
objective comparisons of the performance of different extruder
types have been reported in the literature. Many of the published
evaluations are by extruder manufacturers who are demonstrating
the applications of their particular machines. An exception is an
article by Rauwendaal [33] that indicates that a counter-rotating
extruder is superior for dispersive mixing because of the high shear
stress in the gap. On the other hand, a co-rotating machine gives
better distributive mixing because of the continuous re-orientation
of the shear directions.
The complex flow pattern in a twin screw extruder is also helpful
for devolatilization, because it promotes a high degree of exposure
of fresh surface at a vent port. By use of multiple vent ports it has
been found possible to reduce the solvent concentration of a rubber
slurry from 58% to less than 1% in a single pass through a
counter-rotating intermeshing extruder [34]. In the same work it
was also shown that polyethylene terephthalate scrap and trim film
could be recycled without drying and suffered an intrinsic viscosity
change due to hydrolysis of less than 5%. This remarkable result
was attributed not only to the efficient devolatilization but also to
very efficient melting, which reduces the residence time before
devolatilization can begin.
This last example also illustrates the other advantage claimed for
twin screw extruders, namely the ease of feeding and conveying
difficult materials. The single screw extruder relies on frictional
drag alone to compact and convey the solid feed. Depending on the
details of design, twin screw machines have some degree of positive
displacement action, especially in closely intermeshing counterrotating extruders. The volume contained between the barrel, the
flight flanks, and the bottom of the opposing flight of the second
screw forms a completely closed C-shaped chamber [4, p. 390].
Except for the leakage flow, the material in this chamber is transported forward by the rotation of the screw, independent of fric-

ROLE OF RHEOLOGY IN EXTRUSION

489

tional drag. Normally, twin screw extruders are operated in a


metered mode, i.e. they are starve fed, and the screw flights are
only partially filled, up to the pumping zone. The throughput of the
extruder is thereby decoupled from its pressure generating function.
REFERENCES
1. C. Rauwendaal, Polymer Extrusion, Hanser, New York, 1986.
2. M. J. Stevens, Extruder Principles and Operation, Elsevier Science Publishers,
New York, 1985.
3. Z. Tadmor and I. Klein, Engineering Principles of Plasticating Extrusion,
Van Nostrand, New York, 1970.
4. Z. Tadmor and C. Gogos, Principles of Polymer Processing, John Wiley & Sons,
New York, 1979.
5. R. T. Fenner, in Computational Analysis of Polymer Processing, J. R. A.
Pearson and S. M. Richardson, eds., Elsevier Science Publishers, New York,
1983.
6. L. P. B. M. Janssen, Twin Screw Extrusion, Elsevier Science Publishers,
New York, 1978.
7. M. L. Booy, Polym. Eng. Sci. 21:93 (1981).
8. S. Middleman, Fundamentals of Polymer Processing, McGraw-Hill, New York,
1977, p. 157.
9. J. M. McKelvey, Polymer Processing, John Wiley & Sons, New York, 1962.
10. L. V. Cancio, R. S. Joyner, and P. L. Balin, Plastics Technology 21:40 (1975).
11. M. Dimitrov and R. Hegele, KunststoJfe 61:815 (1971).
12. G. A. Kruder and J. T. Kim, SPE J. 29:49 (1973).
13. D. P. Isherwood, R. N. Pieries and D. Valamonte, Plastics and Rubber
Processing and Applications 4:257 (1984).
14. W. H. Darnell and E. A. J. Mol, SPE J. 12:20 (1956).
15. B. H. Maddock, SPE J. 15:383 (1959).
16. Z. Tadmor, Polym. Eng. Sci. 6:185 (1966).
17. J. T. Lindt, Polym. Eng. Sci. 25:585 (1985).
18. J. F. Ingen Housz and H. E. H. Meijer, Polym. Eng. Sci. 21:352 (1981).
19. C. Rauwendaal, Polym. Eng. Sci. 26:1245 (1986).
20. B. Elbirli, J. T. Lindt, S. R. Gottgetreu and S. M. Baba, Polym. Eng. Sci. 23:86
(1983).
21. J. F. Carley and J. M. McKelvey, Ind. Eng. Chern. 45:985 (1953).
22. C. Rauwendaal, Polym. Eng. Sci. 27:1065 (1987).
23. C. I. Chung, Polym. Eng. Sci. 24:626 (1984).
24. B. H. Maddock, SPE J. 15:983 (1959).
25. D. G. Luenberger, Introduction to Linear and Nonlinear Programming, Addison-Wesley, Reading, Massachusetts, 1973.
26. R. K. Mittal, V. B. Gupta and P. K. Sharma, Composites Sci. & Technology
31:295 (1988).

490

27.
28.
29.
30.
31.
32.
33.
34.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

J. M. Ottino and R. CheBa, Polym. Eng. Sci. 23:357 (1983).


L. ElWin, Polym. Eng. Sci. 18:572 (1978).
L. ElWin and F. Mokhtarian, Polym. Eng. Sci. 23:49 (1983).
J. A. Biesenberger, ed., Devolatilization of Polymers, Hanser Publications,
New York, 1983.
K. Eise, H. Herrmann, S. Jakopin, U. Burkhardt and H. Werner, Adv. Plastics
Tech. 1:18 (1981).
R. J. Nichols, Modem Plastics, Sept. 1986, p. 90.
C. Rauwendaal, Polym. Eng. Sci. 21:1092 (1981).
T. Sakai and N. Hashimoto, SPE (ANTEC) Tech. Papers 32:860 (1986).

Chapter 15
Role of Rheology
in Injection Molding

15.1 INTRODUCTION

In the injection molding process the objective is to produce a


product that is free of voids and sink marks, is not subject to
warpage, and has sufficient strength and stiffness for its end use.
This requires that the melt flow freely into the mold cavity, and that
the final part be reasonably free of residual stresses. At the same
time, the product must be produced at minimum cost, and this
implies the shortest possible cycle time. The challenge, then, is to
produce a good quality part at a minimum cost, and melt rheology
plays a central role in meeting this challenge.
The various areas in which melt flow occurs are shown schematically in Figure 15-1. First, resin is melted, most often in a reciprocating screw extruder. Once a sufficient quantity of melt has been
accumulated, the screw or a ram moves forward and forces the melt
through a nozzle and "sprue" to the "runner" system. The runners
transmit the melt to one or more mold cavities. At the entrance to
the cavity, the melt flows through a "gate," which is a small opening
that facilitates the removal of material solidified in the runners and
minimizes outflow when the pressure is released. In the cavity the
melt comes into contact with the cooled wall and starts to solidify.
Once the mold is full, the continued application of pressure for a
brief time leads to a small additional flow of melt into the cavity
during the "packing stage." This helps to compensate for the
contraction that accompanies the cooling and solidification process.
491

492

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


MOLD

SEPARATES
HERE

MOLD

CAVITIES

Figure 15-1. Sketch of a simple injection mold showing the nozzle, sprue, runners, gates and
mold cavities.
15.2 MELT FLOW IN RUNNERS AND GATES

Runners are normally designed to allow the melt to reach the


cavities while contributing as little as possible to the overall pressure drop and minimizing scrap. Traditionally, the runners were
cooled along with the mold cavity, and the melt therefore solidified
in the runners and was removed as scrap when the mold was
opened to remove the finished part. More recently, there has been
a trend toward the use of heated runners, as this reduces waste and
shortens the cycle time.
A round runner gives the lowest pressure drop for a given flow
rate, but other cross-sections are often used because they facilitate
waste removal and are easier to fabricate than the round design.
The injection pressure required to fill the runners is generally
rather low. However, the pressure drop in a runner can be important in the case of a multicavity mold, because it is highly desirable
to have equal flow rates to all cavities (if the cavities are identical)
and to have all cavities full at the same time. One approach to this
problem is simply to make all the runners of equal length and
diameter. However, the use of such a naturally balanced runner

ROLE OF RHEOLOGY IN INJECTION MOLDING

493

system often involves long runners with high pressure drop and, in
the case of unheated runners, a large amount of regrind.
Some idea of the techniques that can be used to balance the flow
in a runner system can be obtained by the use of equations in
Section 8.2.3 to derive an expression for the flow rate as a function
of pressure drop:

= 7rR 3 (

n
)( -t::.PR )l/n
3n + 1
2KL

(1S-1)

For this simple case of isothermal flow of a power law fluid, the
flow rate for a given pressure decreases with the length, L, and
increases rather sharply with the radius, R. If Q is identified as the
cross sectional area multiplied by the velocity of the melt front in
the runner, Equation IS-1 can be used to calculate the penetration,
L(t), of the melt along a runner as a function of time. The resulting
expression is given by Tadmor and Gogos [1]. It shows that L(t) is
proportional to R and that most of the penetration occurs at the
beginning of the cycle. For a constant flow rate, the pressure should
be a linear function of time.
Of course, for the quantitative design of a runner system the
temperature dependence of the viscosity must be taken into account. For heated, circular runners, Derezinski [2] has presented
dimensionless plots based on the following empirical equation,
which is a combination of a power law shear rate dependence and
an exponential temperature dependence:

.j. )n-l

YJ = YJo ( 'Y 'Yo

-(T-T)/B
0

(IS-2)

The reference values, To, and Yo' are usually evaluated at the wall.
Williams and Lord [3] presented the basic equations for the
velocity and temperature profiles in circular channels. They simulated the transient associated with the start of a filling cycle and
concluded that the steady state result could be used with little
error. Glenn [4] presented a simple method for the preliminary
design of runner systems, and Chen and Hsu [S] have described a
numerical simulation that can be used to balance the flow in a
multicavity mold. The latter authors used an isothermal power law
model in their simulation.

494

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

While the shear rate experienced by the melt in flowing from the
injector system to the gate is usually around 1000 s -1, that in the
gate is much higher. The gate cross section is quite small in order to
facilitate part removal and minimize outflow when the pressure is
released. The nominal wall shear rate in the gate can reach
10 5 s -1, and at this high shear rate the viscosity may be reduced.
Once the viscosity has been reduced by shearing at high rates, it
does not recover instantly when the shear rate drops to a lower
value in the cavity. However, the melt is subjected to the high shear
rate of the gate for only a very brief period of time, and only the
portion of the melt close to the wall experiences the high rate. It is
not clear at this time how much reduction of viscosity actually
occurs in the gate region and how much this influences the cavity
flow near the gate.
15.3 FLOW IN THE MOLD CAVITY

The challenge in filling the cavity. is to achieve complete fiJlin


without short shots while- avoiding sink markl~arp~g(;!~ sticking in
the mold, flash, and poor mechanical properties. This is accomplished by delivering the-correct amount ofresin-to the cavity while
avoiding overpressurization, high thermal stresses, and high residual orientation. Some of the factors that favor complete filling,
however, also promote overpressurization and residual stresses, so
care must be taken in selecting operating conditions for a given
mold and resin.

FLOW FRONT

FROZEN LAYER

FOUNTAIN
FLOW

Figure 15-2. Sketch showing fountain flow phenomenon.

ROLE OF RHEOLOGY IN INJECTION MOLDING

495

As melt flows into the cavity, even for a very simple rectangular
mold the situation cannot be described in terms of pressure flow
between parallel plates with a gap equal to the mold clearance,
because a frozen layer forms immediately at the cavity wall [6].
Moreover, the melt in the center has a lower viscosity than that
adjacent to the frozen layer because of its higher temperature, and
as a result the maximum shear rate occurs not at the interface with
the frozen layer but a short distance away [7]. The maximum shear
rate in the cavity is often in the range of 8,000 to 15,000 S -1.
Another important phenomenon that causes the flow to deviate
from two-dimensional flow between parallel plates is termed the
"fountain effect," first described by Rose [8]. He noted that the
meIU;lQe~ not reach the wall by _a simplejo~
but rather Jends to flow down the center of the cavity to th~lt
irontandthen flow out toward the wall. fhistype-of flow pattern is
illustratecrfn Figure- 15-2: - It can-have a significant effect on the
direction of the flow-induced orientation of the polymer molecules
[9] and thus on the microstructure of the finished part [10, 11].
If the melt must flow around an obstacle in the cavity, such as an
insert or a slot, a "weld line" or "knit line" will result. Once the
melt is separated into two streams, it loses its structural integrity
along the surface of separation, because polymer molecules can
only reestablish a high degree of interlinking across this surface by
a rather slow diffusional process that is driven by thermal (Brownian)
motion. This phenomenon can lead to very large reductions in
tensile strength across the line [13]. Increasing melt temperature
generally increases weld line strength by increasing the intensity of
thermal motion and thus speeding the movement of molecular
segments across the line [14]. The nature of the material being
molded is also an important factor [15].
Pisipati and Baird [16] used interrupted shear to determine a
"reentanglement time," which they hypothesized to be relevant to
the reestablishment of bonds across weld lines. They found this
time to be as much as 100 times greater than the characteristic time
for stress relaxation. They also found that the reentanglement time
increased with molecular weight.
Kim and Suh [17] have suggested that incomplete bonding is only
one of three factors that contribute to weakness along a weld line.

496

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

These factors are:


1. Incomplete bonding at the interface
2. Frozen-in molecular orientation
3. V-notches around the edges of the weld line surface resulting
from gases trapped by the colliding melt fronts in'the center of
a double gated mold.

They concentrated their attention on amorphous polymers, pointing


out that the size and distribution of crystallite structures would also
playa role in the case of a crystalline polymer. They formulated a
model for the first and second factors in the above list and described the behavior predicted by this model. They defined a
"degree of bonding" to serve as a measure of the recovery of full
strength across the weld line. For injection molding grades of
polystyrene, bonding was quite fast as long as the temperature was
above Tg They noted that in practice, it is the time for bonding that
is fixed in the injection molding process, i.e., the time period from
the formation of the weld line until the temperature falls below Tg ,
and that this is a function of the melt temperature. The model
predicted that for a time of 10 s, the degree of bonding is about
50% when the melt temperature is 50C above Tg , which is about
100e. At temperatures more typical of injection molding operations, the bonding was found to be nearly complete at 10 s. They
also used the model to predict the recovery from the molecular
orientation that is generated at the weld line, and this was found to
be the limiting factor at these higher temperatures. Even at these
temperatures, the residual loss in the degree of bonding was only
about 10%. The validity of these conclusions has not yet been
established.
Brewer [18] evaluated the effectiveness of three techniques that
have been proposed to strengthen pieces in the area of a weld line:
1. Heat treatment
2. Solvent exposure
3. Elimination of notch by machining

He feels that tensile tests are not a useful measure of weld line
strength for ductile materials such as ABS, and recommends the

ROLE OF RHEOLOGY IN INJECTION MOLDING

497

Figure 15-3. Sketch of jetting phenomenon.

unnotched Izod impact test. Only solvent exposure was found to


have some effect on the performance of samples in this test.
A phenomenon that can lea~_~~_comJ?JexJ?attern of ~eliLUJ.1es
is "jetting." This term refersto the ten~ency oLth_~ __l!1elt to spurt
into the cavity without wetting the walls near the gate,-ancrthe
result is that the cavity fills by a piling up of the jet at the end of the
cavity rather than by the smooth flow of the melt from the gate.
This phenomenon is illustrated in Figure 15-3. Jetting can occur
when the gate faces a mold wall that is far away and when the flow
rate is high. Yokoi et al. [19] used a glass prism inserted in mold
to observe jetting. They found that in some cases, jetting occurred
in a mold made entirely from metal, but not in a geometrically
identical mold with the glass insert. Oda et al. [20] have suggested
that jetting occurs when the melt swell at the exit from the gate is
insufficient to cause immediate contact between the melt and the
cavity walls. Jetting can be eliminated by redesigning the mold so
that the gate is opposite a nearby wall or by using a fan-shaped gate
or by using a larger gate.
Residual stresses due to nonuniform cooling and residual orientation locked in when the melt cools below its melting point or
below its glass transition temperature are of concern, because they
can result in warpage, delamination, and poor mechanical properties. It is therefore desirable to keep these at a low level. Orientation, which results in anisotropic properties in the finished part, is a
result of the molecular stretching and alignment that accompanies
the melt flow occurring during filling of the cavity [21]. The shear
stresses that are generated during filling provide a rough guide to
the level of orientation that is produced. Thus, a high viscosity and
a high injection pressure usually lead to a high degree of orientation. O-lientati()n_decays with time along with stress as long as the

498

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

resin is in the molten state, but the rate of decay decreases sharply
as the temperature falls. Extensional flow, which IS generated by
.the fountain effect, also produces significant orientation.
Willey and Ulmer [22] suggest that to minimize orientation it is
necessary to minimize the wall shear stress and the thickness of the
solid layer formed during filling. While the first criterion implies a
low filling pressure, i.e., a low filling rate, the latter suggests a fast
extrusion!
In "elongational flow injection molding" [23-25] one makes use
of a flow channel that decreases in area in the direction of flow to
promote orientation in a direction that increases the strength of the
part. Obviously, to use this idea one must have a certain flexibility
in the selection of the mold shape.
Another technique that has been proposed to improve control
over the structure and properties of injection molded parts is the
"multiple live-feed" injection process developed by Allan and Bevis
[26]. In its simplest form, this technique involves the use of two
packing pistons, each feeding a different sprue. During the packing
stage, the two pistons oscillate, rapidly pumping melt back and
forth through the mold during cooling. It is claimed that this
procedure can enhance fiber orientation, reduce weld line weakness
and control microstructure.
Specifying optimal operating conditions for injection molding is
not at all a straightforward operation. The mold temperature must
be below the softening point of the resin, but if it is too low, high
thermal stresses can result in poor part appearance and performance. A low mold temperature will also promote the rapid formation of a frozen wall layer, and this will increase shear stresses and
orientation during filling. Melt temperature has a stron~~Jfect on
both rheological properties and thermal phenomena. For~K~JJl-1l1e,
increasing the temperature reduces viscosity and increases the flow
time before the gate freezes, and this can help to eliminate a
problem with short shots. The higher temperaturell:!_s() leads to
faster relaxation of orientation. On the other hand, an increase in
melt temperature lengthens the cycle time and can lead to sticking
in the mold.
Increasing pressure is another way to achieve faster flow into the
mold. Moreover, as is shown by Equation 15-1, because the melt is
shear thinning the flow goes up with the lin power of the pressure.

ROLE OF RHEOLOGY IN INJECTION MOLDING

499

Since n is less than one for most melts, the flow rate goes up at a
higher than proportional rate with pressure. Some of this gain is
lost because of the increase of viscosity with pressure, but the net
effect will usually be at least a proportional increase of flow rate,
especially in the early stages of the filling process. On the other
hand, increasing the melt pressure requires a higher clamp force
and can also cause sticking, "flash" and high residual stresses,
especially near the gate. Flash is an undesired thin sheet of plastic
attached to the final part that results from the flow of melt beyond
the mold cavity.
The molding of thin parts is a special challenge, because the
viscous resistance to flow is large, and solidification occurs quickly.
Obviously a low viscosity resin is almost essential in this case,
although lowering the viscosity will promote "drooling" from the
nozzle and flash. Drooling can be controlled by the use of a check
valve in the nozzle.
The design of molds to ensure effective filling without the use of
impractically high temperatures or pressures is another challenging
area. Because of the importance of proper mold design, substantial
effort has gone into the development of detailed numerical models
of the filling process. In formulating such a model, a serious
problem arises as a result of the fountain flow phenomenon. There
is a line on the cavity wall at which the melt contacts the frozen
layer, and the necessary motion of this contact line violates the
customary no-slip condition. Kamal et al. [27] have described the
mathematical problems that arise from this situation and have
proposed an empirical procedure for incorporating a slip region in
the neighborhood of the contact line.
A mold filling model must incorporate a rheological constitutive
equation. For models that are designed only to represent the flow
into the mold, purely viscous models are normally used. In such
models the viscosity is a function of temperature and shear rate.
One popular choice is:
T/(T, y) = A exp( - BT)yn-l

(15-3)

However, this model cannot simulate the viscosity at low shear rates
or at temperatures within 100C of Tg For more realistic predictions under these conditions, Philipon et al. [28] have proposed the

500

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

use of an empirical equation that makes use of the hyperbolic


tangent function:
In(7])

In7]o(T) - a(T)tanh[/3(T)ln(y/yo)]

a(T) = 0.45{1 + tanh[ C(T - To)]}

(15-4a)
(15-4b)

The two quantities, In 7]0 and In /3, are assumed to be linear


functions of temperature.
When the simulation is designed not just to model the filling
process but also to predict the microstructure in the final part, a
viscoelastic constitutive equation must be used, except in the core
region where there may be sufficient time at high temperature for
molecular relaxation [29,30].
In addition, thermal properties such as glass transition temperature and thermal diffusivity must be incorporated into the model.
These are governed by the chemical structure of the polymer but
are insensitive to variations in molecular weight distribution. Thus,
for a given polymer, it is not usually necessary to alter these
properties in the model in moving from one grade to another.
Rheological properties, on the other hand, are very sensitive to
molecular weight distribution and can vary greatly from one grade
to another of the same polymer. This factor must be accounted for
in the modelling procedure.
In summary, the detailed modelling of the injection molding
process including fountain flow, solidification and final microstructure, is a very complex problem. Models developed to date are only
applicable to very simple mold geometries.
15.4 LABORATORY EVALUATION OF MOLDING RESINS

There are two approaches to the evaluation of molding resins. The


more fundamental approach is to measure well-defined physical
properties such as viscosity and thermal diffusivity. Such data are
required for a mathematical simulation of the injection molding
process and thus provide a reliable basis for rating the processability of resins. Physical properties are measured under carefully
controlled conditions; for example, rheological measurements are
carried out isothermally.

ROLE OF RHEOLOGY IN INJECTION MOLDING

501

The second approach to the resin evaluation problem involves


nonisothermal tests that simulate, in some sense, the injection
molding process. Such a test does not yield values of well-defined
physical properties, but gives a purely empirical, apparatusdependent measure of processability. These approaches to resin
evaluation are described in the following two sections.
15.4.1 Physical Property Measurement

Both rheological and thermal properties of melts are important in


the injection molding process, and the specific role played by each
depends on the details of the mold design. However, it is more
important in resin grade selection and quality control to measure
rheological rather than thermal properties. The reason for this is
that rheological properties are much more sensitive to molecular
weight and molecular weight distribution than are the thermal
properties. Moreover, the thermal properties are not so strongly
dependent on temperature and shear rate as are rheological properties.
The physical property that is most important with regard to mold
filling is the viscosity. For a given resin, the viscosity depends on
temperature, pressure, and shear rate, and it should be measured
under the appropriate conditions.
Usually, the only rheological information supplied by resin manufacturers is the "flow rate" (ASTM 1238).1 It is explained in
Chapter 8 that this quantity provides a rough indication of the
viscosity at one temperature and at a rather low shear stress. If two
grades, A and B, of a particular linear polymer, say high density
polyethylene, have their melt indexes in a certain order, i.e., MIA>
MI B , implying that TJA < TJB at u :::: 20 kPa, then it is very likely
that their high shear rate viscosities will be in the same order as are
the low shear stress values. However, as is shown by Figure 15-4,
the ratio of the high shear rate viscosities will be much less than the
ratio of the values at a low shear rate. If the shapes of the
molecular weight distribution curves or degrees of long chain
lThe flow rate defined in the standard test method is commonly referred to as the melt
index, but this term is strictly correct only for one of the several extrusion conditions defined
in ASTM 1238.

502

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

y (log

scale)

Figure 15-4. Viscosity versus shear rate curves for two polymers having different molecular
weight distributions. The upper curve is for a material with a broad distribution, while the
lower curve corresponds to a material with a narrow distribution.

branching for the two resins are markedly different, it is even


possible for their viscosity curves to cross, and in this case the
ordering of the two resins would change in going from a low shear
rate to a high shear rate!
Clearly, the most relevant rheological property for flow in runners and mold cavities is the viscosity at shear rates above 1000 S-1
and at temperatures equal to and below the temperature of the
melt entering the runners. The effect of pressure on viscosity
becomes significant in the later stages of the mold filling process.
Furthermore, viscoelastic properties are required for the prediction
of residual orientation.
15.4.2 Moldability Tests

The advantage of empirical moldability tests is that they are simple


to perform and provide direct evidence of the ability of the resin to
fill a mold. The disadvantage is that the shear rate, temperature,
and pressure all vary during the test in an uncontrolled way and the
resulting mold ability index is a complex function of rheological and
thermal properties. As a result, there is no straightforward way to
scale up the results so that they are quantitatively relevant to the

ROLE OF RHEOLOGY IN INJECTION MOLDING

503

filling of any specific mold. In going from the test mold to


the production mold, for example, the relative roles played by the
rheological and thermal properties may be altered. Or, the relative
importance of the roles played by various regions of the viscosityshear rate curve may be changed. As a result, one may obtain a
ranking of several resins that is not a correct indication of the
relative ease with which they can fill a given production mold. The
empirical moldability tests are therefore most useful when used to
compare several resins of the same family, for example several
linear polyethylenes having similar molecular weight distributions
but different average molecular weights.
A widely used moldability test is to inject melt into a standard
mold having a simple geometry involving a long flow path. The
moldability index in this case is simply the "flow length," i.e., the
length of mold filled before freeze-up under standard filling conditions. Discs, spirals, "snakes," and bar molds have all been used in
this way.
The spiral mold is perhaps the most popular of these test molds.
While ASTM standard test method D3123 describes a spiral flow
mold for use with thermosetting molding compounds, there is no
universal standard for thermoplastics. Two commercially available
spiral flow molds are shown in Figure 15-5. The one on the left is
made according to the specifications of the above mentioned ASTM
standard. Using a rectangular spiral die in their evaluations of
polystyrene, polycarbonate and ABS, Yotsutsuji and Komatsubara
[31] found that the flow length was proportional to the injection
pressure. The pressure drop across their gate was negligible.
As might have been anticipated, the single most important physical property that contributes to the flow length is the viscosity at
high shear rates [32], particularly in the range of 1500 to 2500 s - 1
[33], although this range will depend on the resin, the injection
pressure, and the mold dimensions.
Of course the flow length depends also on the thermal properties
of the melt. In his analytical model of the flow in a cooled spiral
mold at fixed wall temperature, Richardson [34] obtained the
following result:
( 15-5)

Figure 15-5. Two spiral flow molds used to evaluate molding resins. [Photo courtesy of Master Unit Die Products, Inc.,
Greenville, MIl.

Gl

C/)
C/)

o
()

:D

"tJ

C/)

c=;

--I

C/)

!;

"tJ

~
:D
or

:l>
Z

:D

01

ROLE OF RHEOLOGY IN INJECTION MOLDING

where:

h
a
K
C

=
=

=
=

505

flow length
smaller dimension of a rectangular spiral mold
thermal diffusivity
power law constant
constant of order 0.1

Experimental results yield values of L corresponding to values of C


between 0.2 and 0.4 [35,36]. Since the thermal diffusivity is much
less sensitive to molecular weight distribution than the viscosity, it is
the high shear rate viscosity that is the governing property.
Hieber et al. [35] carried out a numerical simulation of the flow
in a spiral die and used it together with experimental data to
estimate the four parameters of a temperature and pressuredependent power law model:
1] =

A exp(Ta/T)exp(bP)yn-l

(15-6)

Since there is no universally accepted mold design or set of


molding conditions, the moldability index that is determined in
spiral mold tests has meaning only within a given laboratory. In
fact, it is nearly impossible to specify a truly universal flow length
test, because there are so many variables, and some of these cannot
be satisfactorily controlled [37]. The design of the machinery is
obviously a key factor, and it is not practical to specify, in detail, the
injection molding machine that must be used for the test. For a
given machine and mold design, there are two sets of operating
conditions: machine settings and mold conditions. While mold
conditions are probably more important, the conditions actually
seen by the melt cannot be fixed by merely specifying these. For
example, the temperature of the cavity wall depends not just on the
temperature and flow rate of the mold coolant, but also on the melt
temperature and the injection rate. It is also not practical to fix the
mold pressure. Fritch [37] gives the following advice regarding flow
length testing:
1. Stick to one machine and mold
2. Keep the hydraulic pressure constant
3. Melt temperature is more important than mold temperature,
but a single temperature is not adequate
4. Injection rate influences the shear rate in the mold, and a
single value is not adequate.

506

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Since there is no way to scale up a single spiral mold test to various


specific molding situations, it is not possible to select in advance
"relevant" standard values for the test variables. Therefore, Fritch
recommends that tests be carried out at three melt temperatures
and three injection rates for a total of nine tests. Obviously, the
values selected should cover the ranges of melt temperature and
mold flow rate that are likely to occur in commercial processing.
Another approach to evaluating the processability of a molding
resin is to measure the pressure required to fill a standard mold.
Furches and Kachin [38] have compared the results of several
rheological tests often used to evaluate injection molding resins.
15.5 FORMULATION AND SELECTION OF MOLDING RESINS

For rapid filling at modest melt temperatures and pressures, it is


clearly advantageous that the melt viscosity be as low as possible.
The cycle time is thus short, and thin runners can be used. A low
viscosity is especially advantageous for the molding of parts with
very thin walls. For these reasons, there has been a trend toward
the marketing of molding resins with ever decreasing viscosity. In
spite of its limitations (See Section 15.4) the "flow rate" (melt
index) is commonly used by resin manufacturers to rate the flow
properties of resins, and molding grades with values as high as
100 dgjmin are available.
However, the increase in melt index is usually accomplished by
reducing the molecular weight, and this has the undesired effect of
decreasing the toughness of the finished part, as indicated by its
impact strength. In order to minimize the sacrifice in mechanical
properties for a given application, the molder looks for the lowest
melt index material that will give satisfactory flow in the mold. This
requires the availability of many grades of resin, and this is disadvantageous to the economics of resin manufacturing.
One approach to the problem is to eliminate the highest molecular weight fractions by narrowing the molecular weight distribution.
This leads to the production of so-called "controlled rheology"
grades of resin. Bormuth [39] has discussed the problems involved
in the formulation of controlled rheology polypropylenes.
It has been suggested that narrowing the molecular weight distribution has a negative effect on toughness. However, this effect is

ROLE OF RHEOLOGY IN INJECTION MOLDING

507

not well understood, as the results of carefully controlled experiments have not been reported. In any event, it is possible, at least in
principle, to fine tune the molecular weight distribution to give
good flow and good toughness, as the detailed dependencies of
these two properties on molecular weight distribution are not the
same. This might be accomplished by blending or by the use of a
specially selected polymerization catalyst.
Another approach to this problem is to decouple the melt flow
properties from the solid state mechanical properties. This might be
done by relying on a novel mechanism for building strength in the
final part, perhaps involving the use of fillers with surface treatment, copolymers, blending, crosslinking, or ionomers. There have
been announcements of high flow molding polycarbonates [40] and
polystyrenes [41] that are said to maintain their impact strength,
although the method of accomplishing this was not announced.
The disadvantage of using the melt index to rate molding resins
was noted in Section 10.4, but it is worthwhile to mention this again
in the present context. For example [42], in the case of a series of
polyethylene molding resins, a standard mold filling test revealed
that of two resins having different MWD, the one with a melt index
of 17 had the same mold filling capability as the other, which had a
melt index of 4.
REFERENCES
1. Z. Tadmor and C. G. Gogos, Principles of Polymer Processing, John Wiley &
Sons, New York, 1979.
2. S. J. Derezinski, SPE Tech. Papers 33:245 (1987).
3. G. Williams and H. A. Lord, Poly. Eng. Sci. 15:553 (1975).
4. W. B. Glenn, Jr., Plastics Technology, April 1, 1980, p. 99.
5. S. C. Chen and C. Hsu, SPE Tech. Papers 33:269 (1987).
6. H. Janeschitz-Kriegl, Rheol. Acta 16:327 (1977).
7. H. Van Vijngaarden, J. F. Dijksman, and P. Wesselling, J. Non-Newt. Fl.
Mech. 11:175 (1982).
8. W. Rose, Nature 191:242 (1961).
9. Z. Tadmor, J. Appl. Polym. Sci. 18:1753 (1974).
10. D. McNally, Polym. Plast. Technol. Eng. B(2):101 (1977).
11. M. R. Kamal, E. Chu, P. G. Lafleur and M. E. Ryan, Polym. Eng. Sci. 26:190
(1986).
12. K. Jud, H. H. Kausch and J. G. Williams, J. Materials Sci. 16:204 (1981).
13. R. Boukhili, R. Gauvin and R. Fisa, Plastics Engineering, Nov. 1987, p. 37.

508

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

14. S. C. Malguarnera and A. Manisoli, Poly. Eng. Sci. 10:586 (1981).


15. S. C. Malguarnera and A. I. Manisoli, SPE Tech. Papers 27:775 (1981).
16. R. Pisipati and D. G. Baird, in Polymer Processing and Properties, Edited by
G. Astarita and L. Nicolais, Plenum Press, NY, 1984, p. 215.
17. S.-G. Kim and N. P. Suh, Polym. Eng. Sci. 26:1200 (1986).
18. G. W. Brewer, SPE Tech. Papers 33:252 (1987).
19. H. Yokoi, T. Hayashi, N. Morikita and K. Toda, SPE Tech. Papers 34:329
(1988).
20. K. Oda, J. L. White and E. S. Clark, Polym. Eng. Sci. 16:585 (1976).
21. J. L. White and W. Dietz, Poly. Eng. Sci. 19:1081 (1979).
22. S. J. Willey and A. S. Ulmer, SPE Tech. Papers 32:173 (1986).
23. R. K. Bayer, A. E. Elia and J. C. Seferis, Polym. Eng. Reviews 4:201 (1984).
24. R. K. Bayer, H. G. Zachman, F. J. Balta-Calleja and H. Umbach, Polym. Eng.
Sci. 29:186 (1989).
25. E. Lopez Cab arcos, R. K. Bayer, H. G. Zachman, F. J. Balta-Calleja and
W. Meins, Polym. Eng. Sci. 29:193 (1989).
26. P. S. Allan and M. J. Bevis, Plastics and Rubber Processing and Applications
7(1):3 (1987).
27. M. R. Kamal, S. K. Goyal and E. Chu, AIChEf 34:94 (1988).
28. S. Philipon, J. Villemaire, M. Vincent, J. Agassant, M. de la Lande,
G. Degeneuve and A. Latrobe, SPE Tech. Papers 34:245 (1988).
29. S. K. Goyal, E. Chu and M. R. Kamal, J. Non-Newt. Fl. Mech. 28:373 (1988).
30. H. Mavridis, A. N. Hrymak and J. Vlachopoulos, J. Rheol. 32:639 (1988).
31. A. Yotsutsuji and T. Komatsubara, Kagaku to Koggo 56:263 (1982).
32. I. I. Rubin, Injection Molding, Theory and Practice, John Wiley & Sons, New
York, 1972.
33. J. J. Gouz and G. G. Greygang, SPE Joum., Nov. 1961, p. 1211.
34. S. M. Richardson, Rheol. Acta 24:509 (1985).
35. C. A. Hieber, H. H. Chiang, R. C. Ricketoson, W. R. Jong and K. K. Wang,
SPE Tech. Papers 33:938 (1987).
36. A. M. Hull, S. M. Richardson and J. H. Selopranoto, Plastics & Rubber
Processing and Applications 6:189 (1986).
37. L. W. Fritch, Plastics Engineering 42, no. 6 (June), p. 41 (1986). Also published
in SPE Tech. Papers 32:140 (1986).
38. B. J. Furches and G. A. Kachin, SPE Tech. Papers 35:1663 (1989).
39. H. Bormuth, Kunststoffe 76:428 (1986).
40. "Faster-cycling polycarbonate offered," Chern. & Eng. News, Dec. 12, 1983,

p.5.
41. "No physical property tradeoff in high flow polystyrene resin," Modem Plastics,
Jan. 1986, p. 88.
42. "New high-flow polyolefins make thin-walled molding a better bet," Modem
Plastics, June, 1979, p. 75.

Chapter 16
Role of Rheology in
Blow Molding
16.1 INTRODUCTION

The two principal types of blow molding process are extrusion blow
molding and injection blow molding. In the latter process a "preform," often similar to a test tube with a threaded end, is injection
molded and subsequently reheated and inflated inside a mold. This
process affords excellent control of the thickness distribution in the
preform and is used to make small containers with high quality
finishes. It also lends itself well to the stretch blow molding process
used to make carbonated beverage bottles.
In the extrusion blow molding process, a tube or "parison" of
melt is extruded from a die. The mold halves then close around the
parison to pinch it off at one end and, if a bottle is to be made, to
form a threaded neck at the other. Then the parison is inflated to
conform to the shape of the mold. The extrusion can be intermittent, halting while the parison is pinched, inflated and cooled, or it
can be continuous, by the use of two or more moving molds. The
extrusion process is faster and more economical than the injection
process, and is preferred for manufacturing large products. Very
large items can be made by extrusion blow molding, with parisons
weighing as much as 100 kg or more. For such large objects the
extruder is supplemented by a ram accumulator that permits the
rapid formation of a parison.
The viscoelastic properties of a melt play an important role in
blow molding, particularly in the extrusion blow molding process.
The pressure profiles and flow patterns in the extruder and die are
governed by the viscous properties and the shape of the flow
509

510

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

passages. However, between the time the melt exits the die and the
time it touches the cooled mold walls, it moves and changes its
shape entirely as a viscoelastic response to three stimuli. These are:
1. Molecular orientation generated by flow in the die, which
results in parison swell,
2. The force due to gravity, which causes "drawdown" or "sag"
of the parison, and
3. The blow pressure, which causes parison inflation.
16.2 FLOW IN THE DIE

Approximate methods have been proposed to calculate the important aspects of flow in a blow molding die. It has been found that
the calculation of the major forces and velocity distribution requires
only a knowledge of the viscous properties of the melt, i.e., the
viscosity as a function of shear rate and temperature. While Kim
et al. [1] suggest that the neglect of the temperature change in the
die leads to a large error in the calculated pressure drop in the case
of rigid PVC, models proposed for polyolefin flow generally assume
the flow to be isothermal [2-4].
Such models have been used, for example, to calculate the total
force exerted by the melt on the mandrel [3]. For a converging die,
the normal pressure acts in a direction opposite to the shear force.
This makes it possible to design a die so that the net force on the
mandrel is quite small. This simplifies the die design and makes it
possible to use a stepper motor rather than a servo hydraulic actuator to drive a movable mandrel for purposes of parison programming.
Winter and Fritz [4] used a power law viscosity equation to
calculate the distribution of velocity around the die gap. The
resulting model was then used to design dies having a uniform
velocity distribution, and dies fabricated based on these designs
were found to have very good flow distribution.
Swell is one manifestation of the flow that occurs in the die, but
other effects of die flow include weld lines, shear modification, and
extrudate distortion. If there are "spider legs" to hold the mandrel
in place, weld lines will be formed as the melt flows around the
legs. The motion of molecules to bridge the weld lines and to return

ROLE OF RHEOLOGY IN BLOW MOLDING

511

the melt to a homogeneous state occurs rather slowly, and for this
reason there will be some lateral weakness in the parison and in the
molded product in the neighborhood of a weld line. A second way
in which die flow influences the parison properties is through the
process of "shear modification" or "shear refining" [5,6]. The
effects of this process are most pronounced in highly branched and
high molecular weight linear polyolefins and in blends [6]. The
molecular mechanisms underlying the process are not fully understood, but shearing at high rates alters the structure of the melt,
decreasing the strength of the interactions between molecules, a
process often called "disentanglement." The effect is a reversible
one, and if a shear modified resin is annealed at a temperature
above its melting point, it will recover its preshear structure. However, the time required for this recovery can be much longer than
the time from when a parison is formed to when it is inflated [7].
Thus, the melt that forms the parison and is inflated into the mold
may have been somewhat altered by shear in the die. Shear modification reduces elasticity and can also result in a lower viscosity,
expecially in the case of branched polymers.
Finally, the melt leaving the die may exhibit "sharkskin" or "melt
fracture," which is a distortion of the extrudate that can affect the
surface finish of an extrusion blow molded container. This phenomenon is discussed in some detail in Section 8.9. It occurs above
a critical wall shear stress in the die and is often the factor that
limits the speed of an extrusion process. Extrudate distortion is
most severe in the case of narrow MWD, high viscosity resins. It
can sometimes be eliminated by increasing the temperature or
reducing the extrusion rate, but either of these actions will increase
the cycle time. The use of an internal heater in the mandrel is
thought to reduce melt fracture on the inside of the parison. The
detailed origins of this phenomenon are not fully understood, and
there is some controversy as to whether it can be eliminated by
promoting slip or by eliminating it [8].
Clearly, the cross sectional shape and surface appearance of an
extrudate are governed by many factors in addition to the dimensions of the die lips, and there is currently no reliable method for
predicting these characteristics for a given die, resin and operating
conditions. A further complication is that as soon as melt leaves the
die to form part of a parison, it becomes subject to the force of

512

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

gravity, which leads to sag or drawdown. This tends to make the


parison smaller at the top than at the bottom. The shape of the
parison at the moment of inflation is thus the result of the simultaneous processes of sag and swell.
Because of the complexity of this situation, together with the
complexity of the rheological properties of the melt, it is not
possible to design a blow molding machine to produce prescribed
parison shape and dimensions at the moment of inflation. However,
these dimensions are of crucial importance as they govern the
thickness distribution in the finished product. For this reason machines being used to produce large or irregularly shaped objects are
usually equipped with parison programming devices to alter the
geometry of the die during parison extrusion to produce a parison
having a prescribed shape. However, parison programming cannot
compensate for resin properties that are basically unsuited for use
in a given machine, and for this reason it will be useful to examine
in more detail the processes of parison swell and sag.
16.3 PARISON SWELL

The melt deformation that occurs in the die generates molecular


orientation, which manifests itself at the die exit as extrudate swell,
a phenomenon first described in Section 8.8. Flow at the inlet of a
die, where the streamlines are converging rapidly, involves a high
rate of stretching in the flow direction. This produces a high degree
of molecular orientation, and if the melt were permitted to exit
immediately, as in the case of flow through an orifice plate, there
would be a very high degree of swell. However, if the entrance is
followed by a long straight section, for example a capillary or
straight annular die, molecular relaxation processes will lead to the
loss of the orientation generated at the entrance, and as the die
becomes longer, the degree of swell is reduced. At the same time,
however, shearing in the die produces some axial orientation, and
for a very long die, this will result in a significant degree of swell.
Furthermore, if the die has an expanding or contracting section,
this will produce molecular stretching, which will introduce orientation of a type that depends on the details of the die design. For
these reasons, parison swell is very sensitive to the shape of the die
channel.

ROLE OF RHEOLOGY IN BLOW MOLDING

513

Figure 16-1. Sketch of a parison showing quantities relevant to parison swell.

To describe the swell of a tubular extrudate, it is necessary to use


two swell ratios. For example, the diameter and thickness swells can
be used. These are defined as follows, where the dimensions referred to are defined in Figure 16-1.
(16-1)
(16-2)
It is also possible to define a "weight swell," B w , which is the
weight of a given length of parison, divided by the weight of the
same length of a parison having the same inner and outer radii as
the die. If the density of the melt does not change significantly
while the parison is being formed, the weight swell is equal to the
area swell, BA'

(16-3)

514

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

If the parison wall thickness is small compared to its diameter, the

area swell is approximately equal to the product of the diameter


and thickness swells:

(16-4)
Both the quality and cost of a blow molded container are strongly
dependent on the parison swell ratios. If the diameter swell is too
small, incomplete handles, tabs or other unsymmetrical features
may result. On the other hand, if the diameter swell is too large,
polymer may be trapped in the mold relief or pleating may occur.
Pleating, in turn, can produce webbing in a handle. Weight swell
governs the weight, and thus the material cost, of the molded
product. What is desired is the minimum weight that provides the
required strength and rigidity.
Because swell is a manifestation of the viscoelasticity of the melt,
it is time dependent. For example, for high density polyethylene at
170C, 70 to 80% of the swell occurs during the first few seconds
after the melt leaves the die, while the remainder occurs over a
period of 2 to 3 minutes [9]. For polypropylene at 190C, only 50%
of the swell occurs in the first few seconds, while more than 10
minutes are required to reach an ultimate value [9]. Another source
of time dependency in the case of intermittent extrusion is the
so-called "cuff" effect. The first melt to be extruded after a period
of no flow has had an opportunity to relax in the die and experiences relatively little shear just prior to being extruded. Thus, there
is less molecular orientation, and less swell, than for melt that has
just flowed through the entire length of die. This means that the
swell is least at the bottom of the parison (the "cuff") increasing to
a maximum and then decreasing. Of course sag will cause a decrease in swell toward the top of the parison.
Swell increases as the temperature decreases, although it takes
place somewhat more slowly. It has been suggested that this effect
might be used to control swell by the manipulation of the power for
the die heater [10]. Since the time during which the parison is
exposed to the air before the mold closes is rather short, and the air
surrounding the parison is warm and relatively still, the decrease in
temperature that occurs during parison formation is generally rather

ROLE OF RHEOLOGY IN BLOW MOLDING

515

small, usually less than 5C [11]. For this reason, parison formation
is often assumed to be an isothermal process.
Swell increases as the flow rate increases due to the enhanced
molecular orientation in the die. Swell varies greatly from one
polymer to another and is strongly affected by the extent of branching and the molecular weight distribution. For linear polymers, a
broader MWD generally results in a higher swell. However, resins
with very similar measured molecular weight distributions can have
significantly different swell behavior [12], and this probably reflects
the fact that swell is highly sensitive to small amounts of high
molecular weight material. Koopmans [13] has studied the effects
of small amounts of high molecular weight material on the swell of
HDPE blends at the exit of a capillary at a wall shear rate of
300 s -1. He found that the molecular weight of the high molecular
weight component in his blends was a very important parameter,
and he concluded that it is misleading to use a single parameter
such as Mw/Mn as a measure of molecular weight distribution. The
apparent decrease in swell that Koopmans observed as the molecular weight of the heavier component was further increased may be
due to the time dependency of the swell. Raising the molecular
weight increases the ultimate swell but decreases the rate of approach to this steady state value. Highly branched polymers tend to
swell more, but it is not possible to generalize when both branching
and MWD are altered simultaneously.
Because swell is an elastic recoil process that results from molecular orientation in the die, the shape of the die channel has a strong
effect on both diameter and thickness swells. The simplest case is a
long straight annular die. Here we have only shear flow, and we
would expect to see some orientation in the axial direction. This
suggests that there should not be preferential orientation in either
the radial or hoop direction and that the diameter swell should thus
be equal to the thickness swell.
( 16-5)
This is referred to as "isotropic swell." If Equation 16-4 is valid,
this suggests that
(16-6)

516

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

However, the situation is complicated by the fact that the shear


rate is not uniform across the die gap, but varies from zero, on
some cylindrical surface between the inner and outer walls, to a
maximum at the walls. Cogswell [14] made an analysis based on a
simplified, geometric model of the elastic recoil process and derived
the following relationship:
B H = B2D

(16-7)

If Equation 16-4 is valid, this implies that


BW

B3D

(16-8)

For anisotropic swell, it has been suggested [15] that the following,
empirical, relationship can be used to describe the situation.
( 16-9)

For example, isotropic behavior corresponds to m = 2, while the


Cogswell derivation yields m = 3.
Garcia-Rejon and Dealy [9] found that for HDPE flowing in a
long, straight die, B H was somewhat less than BD at low flow rates,
while isotropic behavior (Eqn. 16-5) was approached at higher flow
rates, i.e., m :::;; 2. For polypropylene, however, they found that the
behavior was intermediate between those described by Equations
16-6 and 16-8, i.e., 2 :::;; M :::;; 3. Geometrically similar straight dies
(same Do/D) produced the same weight swell, at flow rates adjusted to give the same wall shear rate [12].
For more complex dies, such as those used in blow molding
machines, the swell ratios are strongly influenced by two die geometry features, the angle of divergence or convergence of the outer die
wall, and the variation of gap spacing along the flow path. A
diverging die stretches the melt in the hoop direction, and this
should reduce diameter swell by counteracting the axial orientation
generated by the shear flow. Thus, we expect to find that m > 2.
Henze and Wu [15] and Orbey and Dealy [12] found in their studies
of swell from diverging dies that

(16-10)

ROLE OF RHEOLOGY IN BLOW MOLDING

517

or, m =:: 4. The latter authors also found that if the die walls are
parallel, the thickness swell is approximately the same as for a
straight die having the same die lip dimensions.
In the case of a converging die, we have stretching along streamlines and compression in the hoop direction. This will enhance the
diameter swell. Orbey and Dealy [12] found that in a 20 converging
die with parallel walls, the diameter swell was considerably greater
than the thickness swell, giving a value of "m" of about 1.6. In their
studies using a larger converging die with parallel walls, Kalyon and
Kamal [16] found values of m between 1.6 and 1.8. Wilson et al.
[17] also found that their converging die produced a much larger
diameter swell than their straight or diverging dies.
Turning now to the role of a variation in the gap spacing, we note
that many blow molding dies have gaps that are narrowest at the
die lips. This is thought to enhance the stability of the exit flow and
reduce the possibility that the melt will pull away from one of the
die walls before it reaches the exit. This convergence generates an
extensional flow, because of acceleration along streamlines, and
promotes molecular orientation along streamlines, thus enhancing
both diameter and thickness swells. Orbey and Dealy [12] found
that their converging die with a narrowing gap produced extrudate
exhibiting nearly isotropic swell (m = 2).
In the studies referred to above, several different methods were
used to measure swell. The time dependency of swell and the
drawdown due to gravity make this a difficult measurement, and the
different methods do not give the same quantitative results. The
earliest studies of parison swell made use of the "pillow mold" or
pinch-off mold originally proposed by Sheptak and Byer [18]. However, Kalyon et al. [19] used a photographic technique to show that
the pinch-off mold technique can produce unreliable results. The
most reliable method [9,12] is to extrude the parison into an oil
bath having the same temperature and density as the melt and to
take photographs of the parison at frequent intervals. This permits
one to make an accurate determination of swell as a function of
time in the absence of sag. However, this is an elaborate and
time-consuming procedure, and simpler tests based on the use of a
pillow mold continue to find use for the rapid evaluation of commercial resins [20].

518

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Because of the difficulties involved in the use of an actual blow


molding die to determine swell, many attempts have been made to
relate parison swell to more easily measured properties. However,
these have not produced useful correlations. For example, it has
not been possible to establish a relationship between parison swell
and either the viscosity function, 71( y), or the storage modulus
[12,21,22]. The reason for this is that neither of these rheological
properties is an indicator of high-shear-rate viscoelastic effects.
Qualitative relationships have been found between parison swell
and both creep recoil [23] and melt strength [22,24]. However,
these correlations are of use only for the rough ranking of similar
resins. Another approach is to relate parison weight swell to capillary swell, B. Cogswell [14] has derived approximate relationships
between capillary and weight swell, for a straight annular die, and
the recoverable strain. If the recoverable strain is eliminated between the two relationships, we obtain the following relationship
between B wand B:
Bw

0.24

+ 0.73B2

(16-11)

Garcia-Rejon and Dealy [9] compared their experimental data with


this relationship and found that for HDPE, all the data were within
15% of the curve given by Equation 16-11, while their polypropylene data showed weight swells considerably higher than those
predicted by this equation.
The results cited above were based on ultimate swell values
determined by extruding into oil and allowing the parison to come
to a steady state. Kalyon and Kamal [16] have suggested that a more
practical procedure is to consider the parison weight swell that
occurs at short times typical of the parison hold time in the
extrusion blow molding process. They compared this with capillary
swell measured at this same time. They also chose their capillary to
give the same residence time and total strain as the blow molding
die used to measure Bw. They found that their data could be
described by the following relationship:
(16-12)

ROLE OF RHEOLOGY IN BLOW MOLDING

519

Considering the importance of parison swell, the difficulties associated with measuring it, and the unavailability of correlations
between swell and readily measured properties, it would be useful
to be able to predict swell by use of a flow simulation. However, this
has proven to be a very difficult problem, and no reliable models
are currently available. A central problem is the lack of a constitutive equation valid for the complex deformations that occur in the
parison formation process.
Winter and Fischer [25] used Wagner's equation (3-64) to predict
the stress state of the melt at the exit of an annular die, which is
much simpler to calculate than the parison swell. They used the
stress ratio, Ntl2u as a measure of the stress state. From Equation
3-48 or 3-52 we see that this is the recoverable shear at very low
shear rates, although (5-66) shows that it no longer has this significance at high shear rates. Winter and Fischer found that the
predicted value of this ratio is very sensitive to small changes in die
geometry. Kaiser [26] used the Wagner model to calculate the
diameter swell of parisons extruded from dies having complex
geometries and claimed good agreement with measured values.

16.4 PARISON SAG

Sag (drawdown) can cause a large variation in thickness and diameter along the parison, and in an extreme case can cause the parison
to break off. For a Newtonian fluid, sag could be kept under control
simply by using a material with a sufficiently high viscosity. Indeed,
blow molding resins generally have melt index values less than one.
Also, sag becomes more severe as the temperature is increased.
However, since polymer melts are viscoelastic, resistance to sag
cannot always be quantitatively correlated with viscosity.
A number of proposals have been made as to which viscoelastic
property of a melt governs sag. Ajroldi [27] used the linear creep
compliance, as calculated from a tensile relaxation modulus, to
predict sag, while Sebastion and Dearborn [28] suggested that the
extensional stress growth function can be correlated with sag behavior. Dealy and Orbey [29] point out that the sag process is neither a
constant stress nor a constant strain rate process so that neither of
these material functions is directly relevant. For small parisons,

520

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

where the strain rate is small, they propose the use of Lodge's
model (Equation 3-27).
Clearly, reliable methods for predicting sag behavior on the basis
of well-defined rheological properties are still lacking. In their
absence, empirical techniques for evaluating sagging tendency are
employed. Basu et al. [30] made video records of parison length
versus time for this purpose. A simpler procedure that might be
useful is that developed by Swerdlow et al. [31] to determine the
extensibility of film resins. They simply clamp a weight onto the end
of a specified length of extrudate from a melt indexer and measure
the resulting sag during a prescribed time.
When we consider the combined effects of swell and sag, the
situation becomes quite complex from a rheological point of view.
Figure 16-2 is a sketch of parison length versus time curves for
various cases. The first part of the curve reflects the extrusion
period during which the parison is formed. Once extrusion stops,
the length is governed entirely by swell and sag. Curve 1 shows the

::c

l-

e!)

W
....J

oen

a:
~

TIME

Figure 16-2. Parison length versus time for three types of flow behavior.

ROLE OF RHEOLOGY IN BLOW MOLDING

521

case in which there is swell but no sag, while Curve 2 is for the case
of sag with no swell. In the case of an actual parison, Curve 3, there
is an initial recoil followed by a slower increase in length, reflecting
the superposition of swell and drawdown.
A number of models of the parison formation process have been
formulated [28,29,32-34]. All of these models assume that swell
and sag are in some way additive. Unfortunately, none of them can
be used to predict with confidence the behavior of the parison for a
given die, resin, and operating conditions.
16.4.1 Pleating

Pleating (also called "draping" or "curtaining") is a buckling of the


parison that occurs when the melt in the upper portion of the
parison is unable to withstand the compressive hoop stresses due to
the weight of the parison suspended below it. Pleating is often
undesirable, for example causing webbing in handles, but it is
virtually unavoidable for very large parisons.
Obviously, large diameter swell and a small die gap are factors
that will increase the severity of pleating. It is not clear exactly
which rheological properties govern the ability of a melt to resist
pleating, but the viscosity is a rough guide to the "melt stiffness"
that is desired. Thus, high viscosity resins are less likely to pleat, but
increasing the temperature increases the likelihood of pleating.
Schaul et al. [24] proposed the use of the quantity (td/ho11o) as an
indication of the severity of pleating, where td is the parison drop
time and 110 is the viscosity at low shear rate. However, this group
does not account for variations in dia,!l1eter and thickness swell
from one resin to another, and it is most useful for low swell resins.
16.5 PARISON INFLATION

The behavior of the parison during the inflation process is a


manifestation of its extensional flow rheological properties. It has
been observed [35] that the parison does not inflate uniformly and
tends to bulge out in the center, especially in the case of PET resins
[36]. Blow-outs can occur if the ratio of the mold diameter to the
parison diameter (the blow-up ratio) is too high.

522

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The deformation is not the same as that which occurs in the


uniaxial extension experiment described in Chapter 6, but the
results of such an experiment are still thought to be relevant to
inflation performance [28]. In particular, it is thought that extension
thickening (strain hardening) implies that a resin will be easy to
inflate and unlikely to exhibit blow-out, even when the blow-up
ratio and inflation pressure are high. Extension thinning (strain
softening), on the other hand, is thought to imply unstable inflation
and an increased likelihood of blow-outs. This hypothesis is consistent with the observation that low density (branched) polyethylene
is easier to inflate than HDPE. Numerical simulations of the
inflation stage usually make use of rubber elasticity models rather
than viscoelastic liquid constitutive equations [37,38].
16.6 BLOW MOLDING OF ENGINEERING RESINS

In order to take advantage of the economies of the extrusion blow


molding process, there is a growing interest in its use for the
processing of engineering resins. However, many of these materials
were originally developed for injection molding, and they tend to
have properties that make them unsuitable for blow molding.
Specifically they have low viscosity, and if semicrystalline they have
a high melting point and a narrow melting range. Moreover, they
are often highly hygroscopic and sensitive to thermal degradation.
Some of these disadvantages can be overcome by modifications of
the process. Hygroscopic resins must be thoroughly dried immediately prior to processing, and temperatures must be very closely
controlled. Extruder screws having a low compression ratio and a
long transition zone are often advantageous, and it is especially
important to streamline all flow channels to avoid stagnation
regions.
The low viscosity poses a problem at the parison formation stage
because of excessive sag and the inability of the melt to be blown
into deep pockets without tearing. Accumulators are helpful, as
they decouple the extrusion speed from the parison formation
speed and thus reduce the hang time prior to inflation. Methods for
reducing sag include "preblow," which supports the parison by an
upward jet of air prior to inflation, and "prepinch," which involves
a partial inflation of the parison while it is still forming.

ROLE OF RHEOLOGY IN BLOW MOLDING

523

To minimize the equipment modifications required, however, and


to make possible the blow molding of large products, it is highly
desirable to modify the resin to increase its viscosity at low strain
rates. This can be accomplished by altering the molecular weight
distribution, by blending, by use of a comonomer, by cross-linking
or by means of an additive.

16.7 STRETCH BLOW MOLDING

In the stretch blow molding process, the parison is preformed in an


injection molding or extrusion operation and is reheated before
inflation. However, the preform is not heated to a temperature
above its melting point but only to a temperature sufficient to
produce a rubbery consistency. The proper selection of the reheat
time, which governs the inflation temperature, is of central importance in this process. The essential feature of the process is the
generation of a high degree of molecular orientation in the inflation
step. In order to accomplish this it is necessary to ensure that nearly
all of the stretching work is taken up by the polymer as elastic strain
rather than viscous dissipation. This only occurs in a fairly narrow
range of temperatures, between 90 and 115C for PET.
The rubbery state temperature range can be altered by changing
the formulation of the resin. In the case of polypropylene a
comonomer can be used, and in the case of PET, diethylene glycol
can be added. Absorbed water can also act as a plasticizer, reducing
the optimum temperature for orientation [39]. However, such
changes will also affect other characteristics of the resin and can
sometimes result in a deterioration of mechanical properties.
The molecular weight of the polymer also influences the
"processing window" or temperature range for processing [40]. In
the case of PET, the inherent viscosity (LV.) is usually used as an
indicator of molecular weight, and PET resins for bottles normally
have LV. values between 0.72 and 0.85. (See Appendix C for the
definition of the inherent viscosity). Cakmak et al. [41] have studied
the kinematics of the stretch blow molding of PET.
PET is nearly Newtonian above its melting point and is thus
considered unsuitable for extrusion blow molding. However, it has
been suggested [42] that the use of a multifunctional comonomer,

524

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

which produces long chain branching and broadens the MWD, can
render PET suitable for extrusion blow molding.
16.8 MEASUREMENT OF RESIN PROCESSABILITY

The selection of a polymer type for a given blow molding application is governed primarily by functional properties such as gas
permeation, strength, and resistance to solvent and high or low
temperature. Once the polymer has been selected, attention must
be directed toward the processing characteristics, and these are
primarily dependent on rheological properties. It is often possible
to vary the processing performance of a given polymer by means of
additives or by modifying the molecular weight distribution, so that
there is some flexibility in the adaptation of a resin to a given
application. The selection or development of a resin will usually
involve the use of sophisticated rheometers that require the attention of skilled technicians.
Once a resin has been selected and production has commenced,
there is a need for quality control tests to ensure that the resin is
sufficiently uniform from lot to lot, and at least one of these tests
should involve a rheological measurement. However, unlike the
measurements made during a resin selection or development program, a quality control test must be designed so that it can be
carried out economically on a routine basis by relatively unskilled
operators. Because of the important differences between the two
types of measurement, each is discussed in a separate section.
16.8.1 Resin Selection Tests

Because there is such a diversity of blow molding processes and of


shapes and sizes of blow molded articles, it is not possible to specify
a general rheological profile for a blow molding resin. If there is
one single process parameter that is of special importance in a
particular application, for example resistance to parison sag, then
identification of the important rheological property may be straightforward. More commonly, a number of process requirements must
be met simultaneously, and the rheological behavior necessary to
meet one requirement may conflict with that needed to meet
another. Ideally one would design the machinery and tooling to

ROLE OF RHEOLOGY IN BLOW MOLDING

525

match the resin characteristics and optimize the entire process.


However, it is often necessary to use existing equipment and to
remain within specific ranges of operating conditions, and the
problem becomes one of determining rheological characteristics
that can accommodate all of these requirements.
One way to handle such a situation has been described by Schaul
et al. [24]. In this example, only the shot pressure and the die gap
could be adjusted. It was necessary to select a resin that would
simultaneously meet the specifications for parison diameter, bottle
weight, melt fracture, and pleating, but some of these requirements
were in conflict. Decreasing the die gap, for example, increased
parison diameter, but this could lead to unacceptable curtaining.
An understanding of the mechanics of each stage of the process is
useful for defining the effects of the rheological characteristics;
without such an understanding, empirical trial-and-error techniques
are necessary.
It is useful to consider the three principal stages of the blow
molding process, each having its own rheological requirements:
1. Plastication of the resin in an extruder
2. Parison formation
3. Parison inflation.

The extrusion process is described extensively in the literature,


and computer programs for screw and die design are available.
Resin behavior in an extruder is dominated by the dependence of
the viscosity on shear rate and temperature. In principle, one also
needs to know the effect of pressure on viscosity; but this is
usually estimated from an approximate formula rather than being
measured.
Parison formation is the most critical stage of the blow molding
process and also the most difficult one to analyze. The specific
rheological properties that must be considered for a given application depend on the parison geometry and on the machine used for
the process. It is important to note that each rheological property
affects more than one aspect of the parison formation process. For
example, the viscosity in the shear rate range corresponding to flow
through the die governs the pressure required to extrude the
parison. If the extrusion rate is limited by the pressure capability of

526

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the extruder, this limitation will determine the time required to


form the parison, i.e. the "drop time." In turn, if this time is
sufficiently long, excessive sag will occur, and this will affect the
diameter, length, and appearance of the parison. Similarly, if the
high shear rate viscosity is too high, the stress at the desired
extrusion rate may exceed the critical stress for extrudate distortion.
The viscosity at low stress, the drop time, and the parison length
will determine how much sag occurs. The maximum drawdown
stress is initially equal to the total weight of the parison divided by
its cross sectional area, which is equal to (pgL), where p is the
density of the melt, g is the acceleration due to gravity, and L is
the parison length. In principle, the rheological resistance to sag is
a viscoelastic property related to the elastic compliance of the melt
and its relaxation time. If the drop time is long compared to the
relaxation time, viscous flow will be the dominant mechanism, while
at shorter times elastic deformation dominates [11].
Sag is most important in the extrusion of very long parisons. The
drop time is then likely to be long compared to the melt relaxation
time, so that the viscosity contributes more than the elasticity to the
total sag. Furthermore, at very low stresses the extensional viscosity
will be approximately equal to three times the viscosity. If one
restricts the problem to the case of small sag, the case of most
interest, the maximum sag occurs near the middle of the parison
and is given by the quantity (pgLt d/27J), where td is the drop time,
and 7J is an appropriate viscosity.
Extrudate swell is usually the most important property for the
parison extrusion process. It was pointed out in Chapter 8 that
swell cannot be predicted accurately from molecular structure or
from other rheological properties. Furthermore, it depends strongly
on die geometry, extrusion rate, and time after extrusion. It must
therefore be measured directly in deciding on a resin for a given
application or for designing tooling for a given resin. The swell
should be measured under the conditions anticipated in the actual
blow molding process, i.e., at shear rates corresponding to die flow.
If a capillary is to be used, it should have an aspect ratio similar to
that of the blow molding die. Because swell is time dependent, it
may be necessary to use apparatus specially designed for this
purpose. A segmented or flat-parison pinch-off mold might be
attached to a conventional capillary rheometer for example. The

ROLE OF RHEOLOGY IN BLOW MOLDING

527

more closely one can simulate the blow molding conditions the
better.
It should also be noted that die swell as measured with a simple
capillary may not give all of the information required. It is possible
that two polymers with similar capillary swell can have different
values of diameter swell for a given weight swell. In many cases the
discrepancy can be accommodated by adjusting the die gap, but if
this is not possible, or if control of parison diameter is crucial, it
may be necessary to measure both diameter and area swells using
an annular die.
Melt fracture is a problem that is likely to be encountered with
very high viscosity polymers or in high shear rate processes. To a
first approximation one can estimate the likelihood of extrudate
distortion from the die geometry, the flow rate, and the viscosityshear rate curve. As a rule of thumb the critical shear stress for
melt fracture is on the order of 0.1 MPa. If the stress in the die is
estimated to approach this value, direct measurement of the critical
stress by capillary rheometry is suggested.
Extensional flow properties have been touched upon in regard to
sag. Much more important is their role in the parison inflation
process, since this is a rapid extensional flow. Extensional flow
properties govern the parison inflation rate and pressure requirement. Also, their dependence on time and strain rate will affect the
stability of the inflation process. Extension thickening (strain hardening) favors stability, because an initially thin spot is subjected to a
higher than average stress and tends to stretch faster. But if the
resistance to extensional deformation increases with strain, this
increase tends to counteract the increased stress and to allow the
rest of the material to catch up with the thin spot. As has been
discussed in Chapter 6, direct measurement of extensional flow
properties is not a simple task. Recourse to the less rigorous
methods, such as estimation from the entrance pressure drop, may
be satisfactory, especially if there is a background of information on
materials whose processing behavior is known.
To summarize, the minimum information that is needed for
selecting a polymer for a blow molding application is the dependence of viscosity on shear rate at processing temperatures. For
extrusion blow molding, knowledge of the die swell, measured at
the appropriate conditions, is also essential. For very rapid pro-

528

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

cesses and for those in which orientation is important, a measure of


melt viscoelasticity is desirable. Extensional flow properties are
useful for comparing the inflation behavior of various polymers.
Finally, from the viscosity curve it is possible to estimate whether
measurement of the critical stress for melt fracture is necessary.
16.8.2 Quality Control Tests

Once a suitable combination of resin and machine has been selected for a given application, it is necessary to establish routine
tests to assure continued satisfactory operation with a minimum of
machine adjustments. Batch-to-batch variation is a fact of life. It is
the function of a quality control program to ensure that a given lot
of resin does not fall outside the acceptable property range.
Among the requirements for a quality control test are reliability,
speed, and cost, both of the test equipment and its operation.
Above all, the test must relate to a relevant rheological property.
However, it is not essential that a quality control test be sensitive
to only a single rheological property. It may happen that only one of
the properties that affect a given test result fluctuates significantly
during normal resin production. In that case these fluctuations will
correlate with the results of the control test, and the control test,
even though it is affected by more than one property, will still be
useful in monitoring the fluctuations. However, this argument fails
if the resin manufacturer changes the production process, either
deliberately or inadvertently. If such a process change occurs, it is
quite possible that two batches can have identical test results but
perform very differently in processing.
A "flow rate" {melt index} test provides the simplest possible
indication of melt consistency. It is better than nothing but can
yield inconclusive or misleading results, as explained in Chapter 8.
A test that indicates the resistance of a melt to extension and is
simple enough to be considered for quality control purposes is the
"melt strength" test. This test may also be useful for determining
the ease of stripping a parison from a die. It has been found
empirically to relate to an observed difference in the ratio of weight
swell to diameter swell for two polymers that could not be distinguished rheologically by other tests. It should be noted, however,

ROLE OF RHEOLOGY IN BLOW MOLDING

529

that this test is sensitive to many variables, including die swell,


cooling rate, and the temperature dependence of viscosity [43].
REFERENCES
1. H. T. Kim, J. P. Darby and G. F. Wilson, Polym. Eng. Sci. 13:372 (1973).
2. R. A. Worth and J. Parnaby, Trans. Instn. Chem. Engrs. 52:368 (1974).
3. R. J. Pritchatt, J. Parnaby and R. A. Worth, Plastics and Polymers, April 1975,
p.55.
4. H. H. Winter and H. G. Fritz, Polym. Eng. Sci. 26:543 (1986).
5. A. Rudin and H. P. Schreiber, Polym. Eng. Sci. 23:422 (1983).
6. B. Maxwell, E. J. Dormier, F. P. Smith and P. P. Tong, Polym. Eng. Sci.
22:280 (1982).
7. J. M. Dealy and W. Tsang, J. App!. Polym. Sci. 26:1149 (1981).
8. A. V. Ramamurthy, J. Rheol. 30:337 (1986).
9. A. Garcia-Rejon and J. M. Dealy, Polym. Eng. Sci. 22:158 (1982).
10. A. M. Henderson and A. Rudin, J. Appl. Polym. Sci. 31:353 (1986).
11. F. N. Cogswell, P. C. Webb, J. C. Weeks, S. G. Maskell and P. D. R. Rice,
Plastics and Polymers 29:340 (1971).
12. N. Orbey and J. M. Dealy, Polymer Eng. Sci. 24:511 (1984).
13. R. J. Koopmans, Joum. Polym. Sci., Part A 26:1157 (1988).
14. F. N. Cogswell, Plastics and Polymers 38:391 (1970).
15. E. D. Henze and W. C. L. Wu, Polym. Eng. Sci. 13:153 (1973).
16. D. M. Kalyon and M. R. Kamal, Polym. Eng. Sci. 26:508 (1986).
17. N. R. Wilson, M. E. Bentley and B. T. Morgan, SPE Joum. 26, Feb. 1970, p 34.
18. N. Sheptak and C. E. Beyer, SPE Joum. 21:190 (1965).
19. D. Kalyon, V. Tan and M. R. Kamal, Polym. Eng. Sci. 20:773 (1980).
20. L. V. Cancio and R. S. Joyner, Modem Plastics, Jan. 1977, p. 72.
21. J. C. Miller, Trans. Soc. Rheol. 19:341 (1975).
22. C. W. Macosko and J. M. Lorntson, SPE Tech. Papers 19:461 (1973).
23. V. S. Au-Yeung and C. W. Macosko, Modem Plastics, April 1981, p. 84.
24. J. S. Schaul, M. J. Hannon and K. F. Wissbrun, Trans. Soc. Rheo!. 19:351
(1975).
25. H. H. Winter and E. Fischer, Polym. Eng. Sci. 21:366 (1981).
26. H. Kaiser, Xth Int. Congr. Rheol. 2:9 (1988).
27. G. Ajroldi, Polym. Eng. Sci. 18:742 (1978).
28. D. H. Sebastion and J. R. Dearborn, Polym. Eng. Sci. 23:572 (1983).
29. J. M. Dealy and N. Orbey, AIChE J 31:807 (1985).
30. S. Basu, F. Fernandez and C. Rauwendaal, SPE Tech. Papers 28:723 (1982).
31. M. Swerdlow, F. N. Cogswell and N. Krul, Plastics and Rubber Processing,
(1980).
32. M. R. Kamal, V. Tan and D. Kalyon, Polym. Eng. Sci. 21:331 (1981).
33. A. Dutta and M. E. Ryan, J. Non-Newt. Fl. Mech. 10:235 (1982).

530

34.
35.
36.
37.
38.
39.
40.
41.
42.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

S. Basu and F. Fernandez, Adv. Polym. Technol. 3, no. 2, p. 163 (1983).


M. E. Ryan and A. Dutta, Polym. Eng. Sci. 22:569 (1982).
L. Erwin, M. A. Pollack and H. Gonsalez, Polym. Eng. Sci. 23: 826 (1983).
c. A. Taylor, SPE Tech. Papers 35:931 (1989).
W. P. Haessley and M. E. Ryan, SPE Tech. Papers 35:934 (1989).
S. A. labarin and E. A. Lofgren, Polym. Eng. Sci. 26:620 (1986).
R. L. Fifer, SPE Tech. Papers 27:696 (1981).
M. Cakmak, 1. L. White and 1. E. Spruiell, J. Appl. Polym. Sci. 30:3679 (1985).
R. Edelman, F. M. Berardinelli and K. F. Wissbrun, U.S. Patents: 4,161,579
(July 17,1979); 4,219,527 (Aug. 26 1980); 4,234,708 (Nov. 18, 1980).
43. K. F. Wissbrun, Polym. Eng. Sci. 13:342 (1973).

Chapter 17
Role of Rheology in Film
Blowing and Sheet Extrusion
17.1 THE FILM BLOWING PROCESS
17.1.1 Description of the Process

The essential elements of the film blowing process are illustrated in


Figure 17-1. An extruder melts the resin and forces it through a
screen pack and an annular die. The extruded melt, in the form of a
tube, flows upward under the influence of a vertical, "machine
direction" force, applied by means of nip rolls some distance above
the die. There is an overall stretching of the polymer in the
machine direction, and the ratio of the linear speed of the film
through the nip rolls, divided by the average melt velocity at the die
lips, is called the "draw down ratio" (DDR).
The tube is cooled by means of an "air ring" that directs air
along its outer surface. In some cases additional cooling is provided
by an internal cooling device. When the process is started up, air is
introduced through a hole in the die face to prevent the collapse of
the tube of molten polymer. After the tube is threaded between the
nip rolls to form an air seal, additional air is introduced, and at
some level above the die lips the tube inflates in a radial direction
to form a "bubble." Here the tube expands to a larger diameter,
and the ratio of this diameter to that of the die lips is called the
"blow up ratio" (BUR). This inflation process generates stretching
in the circumferential or "transverse" direction. Just above the
level at which the larger diameter is established is the "frost line,"
where solidification occurs. This terminology arises from the loss of
transparency due to crystallization in polyethylene film and not
531

532

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

GUIDE ROLLS

FILM
BUBBLE
FROST LINE

MANDREL
EXTRUDER

/'"

r
AIR

HEATERS

Figure 17-1. Schematic diagram of the film blowing process. The air ring (cooling system) is
not shown.

from the presence of any condensed moisture on the film. Also, it is


not a true "line" but a narrow zone over which solidification occurs.
In fact, the solidification occurs first at the cooled surface, and the
melt-solid interface then moves away from this surface through the
thickness of the film as it moves upwards, as shown in Figure 17-2.
The enlarged tube continues its upward movement until it is
sufficiently cool to be flattened without sticking to itself or to the
apparatus. The tube then passes through a frame or flattening
ladder to be reshaped so that it can pass between the nip rolls that
generate the vertical (machine direction) tension. The flattened
tube is then wound up and cut for shipment or for further processing. The width of the flattened tube is called the "layfiat width,"
and this is equal to 17" times the final bubble radius.

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

533

MELT-SOLID
INTERFACE
SOLIDIFIED
LAYER

INTERIOR
OF BUBBLE

OUTER BUBBLE
SURFACE

Figure 17-2. Cross section of bubble wall in region where solidification is occurring. Only the
outer surface is being cooled.

Rheological properties play an important role in film blowing.


They govern the shape and stability of the bubble and the onset of
sharkskin (surface roughness). Because of the complexity of the
flows involved, it is not generally possible to establish simple quantitative correlations between these phenomena and easily-measured
rheological properties. However, an understanding of how variations in the rheological behavior of melts can affect the processing
and properties of blown film is essential if one is to achieve
optimum results from this process.
17.1.2 Criteria for Successful Processing

The objective of the process is to produce a thin film having a


uniform gauge and good optical and mechanical properties. Since
the film is quite thin, it is especially important to avoid the presence
in the extrudate of unmelted material, gels or foreign matter, as
these will be readily apparent in the finished product. In order to
achieve good mechanical properties it is often advantageous, particularly in the case of packaging films, to have molecular orientation
in the film that is more or less equal ("balanced") in the machine
and transverse directions.

534

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The processor wishes to minimize costs, and this implies a high


production rate. To produce a thin film without the processing
difficulties that would result from the use of very narrow die lips, a
high degree of melt stretching is required, and this is achieved by
use of a large DDR value. To minimize the diameter of the die
required to produce a film of given layflat width, a large blow up
ratio is desirable, and values of BUR up to three are common. A
high value of the BUR also helps to compensate for the tendency of
the large DDR to generate high orientation in the machine direction and thus an unbalanced film.
17.1.3 Principal Problems Arising in Film Blowing

An imbalance in the molecular orientation in the finished film can


result in reduced impact strength and the production of a "splitty"
film that tends to tear rather easily in the machine direction.
Molecular orientation is governed by the strains occurring in the
bubble and the rheological properties of the polymer.
A high production rate implies a high cooling rate. If one tries to
accomplish this by simply increasing the air flow to the air ring,
bubble vibration or instability will occur, leading to nonuniform
gauge or even rupture of the bubble. A higher cooling rate can be
achieved by the use of refrigerated air and internal bubble cooling,
but these solutions increase capital and operating costs. The cooling
rate can also be increased by use of an air ring specially designed to
minimize the aerodynamic disturbances to the bubble.
Gauge non uniformity is a serious defect that can lead to the
rejection of finished film. As mentioned above, it can result from
bubble instability, but there are other possible causes, including
extruder surging.
Another important problem in film manufacture is the occurrence of "sharkskin" on the surface of the film. This results from
the flow at the exit of the die, and it causes a serious deterioration
in the optical properties of the film.
17.1.4 Resins used for Blown Film

The resins most used for blown film are polyolefins, particularly
polyethylene and polypropylene. At the same time there is a grow-

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

535

ing use of other resins and of the coextrusion process to make


multi-layer films.
Traditionally, the highest volume resin for films has been low
density polyethylene (LDPE). This material exhibits good bubble
stability and produces a film that is suitable for many applications.
For this material, it has been found that a narrow molecular weight
distribution and a modest degree of long chain branching give good
drawdown and optical properties, although this combination of
characteristics does not give the best impact strength.
Where a stronger, less flexible film is needed, high-molecular
weight, high density polyethylene (HDPE) has been the popular
choice, although it is more difficult to achieve a good balance of
orientation in the machine and transverse directions, probably
because of its extensional flow properties (see Chapter 6).
Starting in the 1980s, there has been a substantial replacement of
low density polyethylene, produced by means of a high pressure
process, by "linear low density polyethylene" (LLDPE), which is a
copolymer of ethylene with a higher alpha-olefin, often butene,
hexene or octene. This resin has the same low level of crystallinity
as the LDPE produced by the high pressure resin, but it can be
manufactured at modest pressures, thus reducing capital and operating costs. Furthermore, it can be more easily drawn down than
LDPE, and this makes it possible to produce a thinner film having
the same strength as one made from LDPE, or a stronger film
having the same thickness. However, there are also some problems
involved in converting production from the one resin to the other.
Most of the LLDPE now produced has a high molecular weight
and a relatively narrow MWD. At the same time, its short side
branches do not have a significant effect on melt rheology. Therefore, compared with LDPE, which has a high degree of long chain
branching, a typical film grade LLDPE has a higher viscosity at the
high shear rates occurring in the extruder and the die, thus increasing screw torque, barrel wear, and melt temperature, and increasing
the likelihood of sharkskin at the die exit. This is illustrated in
Figure 17-3, which shows typical viscosity versus shear rate curves
for LDPE and LLDPE film resins. At the same time, LLDPE
exhibits lower extensional stresses at the low strain rates occurring
in the molten tube and bubble inflation regions, and this renders
the bubble less stable.

536

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

log(y)

Figure 17-3. Typical viscosity-shear rate curves for LDPE and LLDPE film resins.

17.2 FLOW IN THE EXTRUDER AND DIE; EXTRUDATE SWELL

The extrusion of LD PE is well understood, and screw design is a


well established art. In the case of HDPE film resins, special
problems are encountered, as simple screw designs tend to produce
inadequate solids conveying at moderate and high screw speeds [1].
Thus, as screw speed increases, output rate and film properties
deteriorate. Mixing sections can correct this, but their use leads to
higher melt temperatures and thus to degradation. This in turn
reduces the maximum blow up ratio, which is needed to promote
balanced orientation. Grooved barrels in the feed zone can improve
solids conveying, as can barrier screws, which keep unmelted solids
separate from the melt and pushed up against the wall. This
generates less shear and reduces melt temperature and pressure
fluctuations.
The use of LLDPE also poses special problems in extrusion, as
compared to LDPE. The use of barrier screws for LLDPE has been
discussed by Christiano [2]. Because of the higher viscosity at typical
extrusion shear rates, as shown in Figure 17-3, there is a higher
torque on the screw, accelerated barrel wear, and a higher melt

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

537

Figure 17-4. Partial cross section of spiral mandrel die, showing both the main flow path
along the channel and the leakage flow over the land into the next channel. Adapted from
Ref. 5. Copyright 1972 by The Society of Plastics Engineers. Reprinted by permission.

temperature [3]. Solutions to this problem include the use of a


shorter screw or the use of a variable pitch screw [4].
In the area of die design, the objective is to produce an annular
extrudate that has no weld lines and is uniform in thickness and
velocity around its circumference. The most popular type of die for
film is the spiral mandrel die. Such a die is shown in partial cross
section in Figure 17-4. The melt enters from the extruder into the
center and flows out through several feeding ports to a series of
helical grooved channels cut into the wall of the mandrel. The cross
section of these channels decreases in the flow direction, while
there is an increasing gap between the "land" of the channels and
the cylindrical outer wall of the die. Thus, an increasing quantity of

538

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

the melt "leaks" over this land into the next channel, and there is
thus substantial mixing. The depth of the helical channels ultimately reduces to zero, and the final section of the flow channel is
an annulus whose gap varies in the axial direction, narrowing
significantly near the exit.
The flow is a rather complex one, but a very simplified method of
calculation was used by Proctor [5] to calculate the details of the
flow for a purely viscous material with a power law viscosity. More
elaborate methods for power law fluids that make fewer simplifying
flow assumptions and require more computation were described
later [6-10], and several of these have been compared by
Perdikouliaset al. [11]. A model that incorporates viscoelastic material behavior has been developed by Kalyon et al. [12].
Gates [13] has discussed the problems involved in designing a
spiral mandrel die for use with LLDPE and for maximum versatility. He says that the main problem is to avoid the generation of high
shear rates, which result in high back pressure and melt temperature. Furthermore, he points out that there can be a substantial
temperature variation across the diameter of the extruder adapter
that feeds the die, and that the flow in the spiral mandrel die tends
to produce a layering of polymer having different temperatures in
the molten tube finally produced. This in turn can cause cyclic
variations in gauge and frost line height. He suggests that this effect
can be minimized by designing the die to have a longer spiral wrap
distance, leading to more port overlap. A longer spiral channel also
yields performance that is less sensitive to changes in the viscosity
of the resin.
In Section 8.9 we described the types of extrudate distortion that
can occur when melt flows out of the end of a die so that a free
surface is formed. In the film blowing process, the most troublesome of these is sharkskin, also known as surface melt fracture,
because it causes the finished film to have a rough, unattractive
surface. This phenomenon has been much studied over a period of
many years, but its origin is still not fully understood. It appears to
be a surface phenomenon that results from the abrupt change in
the boundary condition from no-slip (in the die) to a free surface.
Cogswell [14] has proposed that this process generates a high
degree of stretching in the surface layer of the polymer as it exits
the die, and that this can produce stresses that exceed the strength

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

539

of the melt, leading to a surface fracture. Kurtz [15] has suggested


that the necessary conditions for the occurrence of sharkskin involve both a critical wall shear stress in the die and a critical stretch
rate at the exit.
Sharkskin is especially troublesome in the processing of LLDPE,
with the result that dies designed for LDPE can only be used at
significantly reduced production rates. The most common solution
to this problem is the use of a wider die gap to reduce the wall
shear rate. However, this increases the demands on the bubble
cooling system and causes some deterioration of film properties.
The use of a die having a diverging gap at its exit to reduce the
shear rate [15, 16], and of heated die lips [17], to reduce the
viscosity and thus the shear stress, have also been proposed. Another approach to this problem that does not require equipment
modification is the use of "processing aids." These are usually
fluorocarbon elastomers [18,19], that can be added to the resin at
time of processing or introduced as a masterbatch. Here capital
costs associated with resin conversion are replaced by increased
operating costs.
Ramamurthy [20] has carried out an extensive study of this
phenomenon. He concluded that surface melt fracture results from
a loss of adhesion between the melt and the wall of the die, and
that this occurs at a critical wall shear stress between 0.1 and 0.14
MPa for all polyethylenes, regardless of molecular structure and
temperature. He hypothesized that fluorocarbons are effective because they promote adhesion. Since this effect requires the build up
of a layer of additive on the die wall, it only appears after additivecontaining melt has flowed through the die for a significant "induction time," which is on the order of an hour. For this reason, the
effect may not show up in tests involving a capillary rheometer
rather than an extruder-fed die. Ramamurthy's hypothesis as to the
mechanism of action of processing aids contradicts other proposals
[18,19] that the additive promotes slip rather than adhesion, as
indicated by pressure drop measurements. Ramamurthy [20-23] has
proposed that sharkskin can be eliminated without a reduction in
flow rate by the judicious choice of the material of construction of
the die combined with the use of an additive.
It was mentioned in Section 5.3.3 that extensive shearing at high
rates, as in an extruder, alters the rheological behavior of a melt,

540

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

especially in the case of polymers with a high degree of long chain


branching. While this effect is reversible, the polymer must either
be dissolved in solvent or held above its melting temperature for a
long period of time in order to recover its pre-shear properties.
Rokudai et al. [24] reported that extensive preshearing reduces
haze and improves gloss in LDPE blown film. To achieve a substantial improvement, however, the resin had to be extruded five times.
Mtinstedt [25] has pointed out that the only difference between the
much studied LDPE film resins IUPAC A and IUPAC C was that
IUPAC C had been subjected to "mechanical pretreatment." He
found that this preshearing had a marked effect on the extensional
viscosity. He also found that extensional flow is more effective than
shear as a mechanical pretreatment. Paradoxically, the recoverable
strain was little affected by pretreatment.
These results suggest that if one wishes to use a rheological test
to evaluate the processing behavior of a highly branched film resin,
the samples should be subjected to preshearing to bring them to a
structural state equal to what it would be at the die exit in the film
blowing process.
17.3 MELT FLOW IN THE BUBBLE

The melt leaving the die is suddenly free of the restraints imposed
by contact with the die wall and it reacts quickly to its new
environment. In particular, it swells in response to the molecular
orientation induced by the flow in the die, and it stretches in the
machine direction in response to the tension imposed by the nip
rolls. Its behavior in such a free surface flow is strongly dependent
on its rheological properties, with thermal properties also playing
an important role. The design of the die and the film cooling system
are also of obvious importance.
For a given resin and film line, the processing conditions that
influence the flow in the bubble include:
1.
2.
3.
4.
5.

Melt temperature
Output rate (melt flow rate)
Nip roll speed
Internal bubble pressure
Cooling air temperature and flow rate.

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

541

The draw down ratio and blow up ratio are usually specified in
place of the nip roll speed and bubble pressure, respectively.
Obviously there are many variables that contribute to bubble behavior, and the systematic study of the effects of all of these is an
enormous task. In fact no such study has ever been carried out, and
our present understanding of the process is based on inferences
from a number of small-scale studies in which only a few variables
were examined at a time. Furthermore, there is no generally accepted procedure for scaling results obtained with one apparatus so
that they are relevant to melt behavior in an apparatus of a
different size and having different design features. Therefore, in
this section, we present some empirical observations together with
some proposed generalizations, but it is important to keep in mind
that these latter do not constitute fundamental principles at this
time.
17.3.1 Forces Acting on the Bubble

The bubble is supported from above by the nip rolls, and its lateral
position is stabilized to some extent by various types of guides,
including irises and cages and by the collapsing ladder. However, at
its base, the bubble is held in place only by the tube of molten
polymer coming from the die. This allows the bubble to move and
deform in response to rather moderate external forces. In particular, the aerodynamic forces arising from the interaction between
the cooling air and the bubble surface can displace and deform the
bubble. This, in turn, can cause gauge variations and even bubble
rupture. Therefore, it is of importance to understand all the forces
that can act on the bubble.
Looking first at the entire bubble, we see an upward (machine
direction) force exerted by the nip rolls. Bubble stabilizing devices
and the collapsing frame will also exert forces on the bubble. At the
bottom, viscous stresses act in the molten tube at the exit of the die
to hold the bubble in place. Internal pressure supplies the basic
driving force for inflation, while aerodynamic forces can operate to
aid or to deter inflation, to cause motion of the entire bubble, or to
stabilize the position of the bubble. Finally, gravity acts directly on
the bubble, so that the nip rolls not only produce the force that
generates stretching in the bubble but must also support its weight.

542

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 17-5. Stresses acting on an element of melt in the expanding region of the bubble. lIM
is the normal stress in the machine direction, and lIT is the normal stress in the transverse
direction.

To gain insight into the way that the operating variables influence
bubble shape, it is useful to take as a system for analysis a small
element of the bubble, as shown in Figure 17-5. The forces acting
on this element are as follows:
Gravity: This acts in a downward direction.
Pressure: This is the difference between the internal and external

pressures on the bubble surface. The contributing factors are the


internal bubble pressure and the aerodynamic forces resulting
from the flow of the cooling air.

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

543

Viscous stresses: The components that act to stretch the element


are those in the machine and transverse directions. This stretching is resisted by the viscoelastic response of the melt.
The shape of the bubble, its motion, and the molecular orientation in the finished film are governed by the interaction between
these forces and by the temperature distribution.
17.3. 1. 1 Viscous stresses in the molten region of the bubble

The principal strain rates in the machine, x M ' and transverse, x T '
directions are easily written in terms of the velocity, vM(t), and
bubble radius, R(t), of a moving element of fluid in the bubble.
However, it is more convenient for describing experiments and for
the modelling process to use dependent variables such as vM(z) and
R(z) that refer to a point in the bubble that is at a fixed distance, z,
above the die. Assuming axial symmetry, the strain rates for both
frames of reference are given below.
(17-1)
.
ET =

dR(t)

R(t) ~

vAz) dR(z)
R(z)
dz

(17-2)

The stresses generated in each of these directions will be related to


the strain rates through a viscoelastic constitutive relation for the
particular polymer involved. Since neither the strain rate nor the
total strain is sufficiently small for the linear theory of viscoelasticity
to be valid, each stress component will depend on the histories of
both strain rates. In selecting a material function from among those
defined in Chapters 2 to 6 that might be useful for correlation with
processability, the several extensional stress growth coefficients
appear to be the most relevant. In selecting from among these, we
note that the deformation occurring in the bubble lies somewhere
between (symmetrical) biaxial extension and planar extension. For
long stalk bubbles the deformation in the tube of molten polymer
below the inflation region is very close to planar extension. Existing

544

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

processability correlations, however, are based on uniaxial extension, which is in fact the least similar to the deformation that occurs
in a film bubble. This is because, of the three types of extensional
flow mentioned, uniaxial extension is the easiest to generate in the
laboratory.
Detailed strain histories in the molten zone of low density
polyethylene film bubbles have been determined experimentally
[26, 27]. The machine direction strain rate is highest at some point
above the die and then decreases monotonically up to the frost line.
The strain history in the immediate neighborhood of the die lips is
complicated by the abrupt change from pressure-driven shear flow
in the die to free surface flow just above it, which involves some
degree of extrudate swell. The transverse strain rate only becomes
significant in the inflation zone where it reaches a maximum and
then falls to zero at the frost line. For some operating conditions,
the transverse strain rate can exceed the machine direction strain
rate in the region of maximum rate of increase of the bubble radius.
However, even in this case, the resulting film may still have a
somewhat preferred orientation in the machine direction [26]. One
way to describe the varying strain history of a material element as it
moves upward as part of a bubble is to plot its trajectory on a strain
invariant diagram such as that shown in Figure 3-l.
Because of the complexity of the flow field and the non-isothermality of the film blowing process, there are no simple cause-effect
relationships that can be used to predict the process performance
of a resin on the basis of a few easily performed laboratory tests.
The complexities resulting from the interaction between the temperature distribution and the rheological properties have been
explored by Gupta et al. [28].
17.3. 1.2 Aerodynamic forces

The general nature of aerodynamic forces can be understood by


reference to Bernoulli's equation, which applies to any point within
the flowing medium and can be written as follows.
dv 2

dP

+gdz+ -

=0

(17-3)

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

545

where p is the density of the air. This equation relates changes in


the pressure (P), elevation (z) and velocity (v) of the air along a
given streamline. It is rigorously valid only for the flow of an
inviscid, incompressible ("ideal") fluid. However, because of the
low viscosity of air and the relatively small pressure changes occurring in cooling air streams, it is a good approximation to the actual
case, except very near to the surface of the film. It can be used as a
guide to the average forces exerted by air flowing near the surface,
although it does not account for turbulence, which causes buffeting
of the bubble.
In the case of cooling air flow, the gravity term is much smaller
than the pressure and acceleration terms, and Bernoulli's equation
can be written as:
2

t1.P
+-=0

(17-4)

Thus, an increase in velocity is associated with a decrease in


pressure and vice versa. Consider, for example, stagnation flow in
which air flows directly toward a solid surface as shown in Figure
17-6. The velocity at the stagnation point is zero, and if the velocity
and pressure far from the surface are VI and PI' the pressure P2 at
the stagnation point is related to these quantities by:
V~

~,P,

(PI - P 2 )
p

(17-5)

Figure 17-6. Stagnation flow. P 2 is the stagnation pressure, and PI is the lower pressure at
some distance away from the wall.

546

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

V"P,

Figure 17-7. Venturi flow. P2 is the pressure at the "throat"; it is lower than Pl.

Thus, the stagnation pressure, P 2 , is higher than PI and is given by:

(17-6)
Another flow situation that is relevant to blown film cooling is
venturi flow, in which the air flows through a contraction that
causes an increase in velocity, as shown in Figure 17-7. To infer
something about the pressure, P 2 on the surface, we make use of
the observation that the pressure gradient normal to the streamlines is very small. Thus, we can use Equation 17-4 to show that:
(17-7)
Since V 2 > VI' we conclude that P 2 < Pl'
These simple concepts show that the pressure on the outside
surface of the bubble can vary substantially from point to point and
can differ markedly from the ambient pressure. Cao and Campbell
[29] have carried out a more detailed analysis of air ring aerodynamics and have compared computational results with pressure
measurements made on a solid model of a film bubble. For a
simple, single-lip air ring that is assumed to generate a vertical,
tangential wall jet, the pressure is found to be positive at a concave
surface and negative at a convex surface.

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

547

Regarding the heat transfer rate, we note that it is much higher


for stagnation flow than for parallel flow. Early air rings were
designed to maximize cooling rates and the air impinged directly on
the bubble surface, generating a stagnation flow. However, this
generated a high positive pressure as well as severe buffeting. To
avoid the destabilizing effects of this flow pattern, air flow rates
were limited to rather low values. This led to the use of a deflector
that diverted the air so that a "fan spray" flow pattern was generated. This permitted an increase in the air flow, but the flow
passages were not streamlined, and large scale turbulence was
generated.
Although this type of air ring was widely used for several decades,
efforts continued to develop improved units. Corbett [30] proposed
the use of the venturi effect to create a smoother air flow and avoid
stagnation flow forces. This improved bubble stability, but the
suction resulting from the venturi effect tended to draw the melt
toward the air ring, and this effect became the limiting factor for air
flow rate. Later, multiple, stacked air rings [31,32] were proposed
to solve the problem of generating streamlined flow without generating large suction at the base of the bubble.
In the case of linear low density polyethylene, the bubble is
particularly sensitive to destabilizing forces, and the venturi effect
has been used to control bubble shape while stabilizing the position
of the bubble. The "dual lip" air ring [32] has a lower lip that
provides initial cooling without suction and a second lip designed
to generate suction to draw the molten tube out and promote
inflation.
The direct measurement of the temperature of the melt in the
bubble is difficult, because the infrared thermometers generally
used respond mainly to the temperature of the outer surface.
However, it is known that very high rates of temperature decrease
do occur [27,33].
17.3.2 Bubble Shape

The two bubble shapes most often observed in commercial film


blowing are shown in Figure 17-8. The "pocket" shape shown on
the left is used for LDPE, LLDPE, and polypropylene, as well as a
number of other materials, while the "long stalk" shape shown on

548

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure 17-8. Typical bubble shapes for LDPE and LLDPE (on left) and for HDPE (on right>.
The air ring is not shown.

the right is used primarily for HDPE. The bubble shape plays an
important role in generating molecular orientation, since the transverse orientation is a direct result of the radial expansion that
accompanies the inflation stage of the operation.
The long stalk process is found to be necessary for HDPE in
order to produce film having a balanced orientation. First, it provides time for the relaxation of the molecular orientation generated
by flow in the die. In addition, we note that in this type of
operation, most of the machine direction drawdown occurs before
inflation, while the transverse stretching occurs at a later time and
at lower temperatures. In this way, the transverse extension makes
a significant contribution to film orientation even though it is
smaller in total magnitude than the machine direction drawdown.
Another reason for using a long stalk, and thus delaying inflation
with HDPE, is that the melt is highly susceptible to instability and
rupture at the high temperature of the extrudate.
Small variations in resin characteristics, for example those due to
batch to batch differences, can cause significant variations in bubble
shape [34], and these differences are amplified by the bubble
cooling process, which is itself very sensitive to bubble shape.

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

549

17.3.3 Drawability

An important feature of the film blowing process is its ability to


generate a high degree of drawdown, so that the gauge of the final
film is much less than the die gap. However, there is a limit to this
process, and this is imposed by the strength of the melt in the
bubble. For a given die, blow up ratio, and throughput rate, if
the nip roll speed is increased, a point will be reached at which the
tensile stress in the bubble exceeds the cohesive strength of the
melt, and rupture will occur. The ability of a resin to be drawn
down can be characterized by the maximum drawdown ratio (DDR),
or by the ratio of the die opening to the film thickness (DDRXBUR).
This limitation is particularly noticeable in the case of LDPE,
and it is thought that this is related to the fact that in uniaxial
extension experiments, this resin is extension thickening (strain
hardening). This means that its tensile stress growth coefficient
increases sharply above the linear viscoelastic curve as the strain
grows. If the principal stretching stresses in the film blowing process
also increase sharply with strain, then such a resin would be
particularly subject to cohesive failure. However, it must be remembered that flow in the bubble is not a uniaxial deformation, and the
correlation between extension thickening and the tendency to rupture is speculative at the present time.
In a comparison of three similar LDPE film resins [34], it was
found that differences in maximum film drawdown had little relationship with rheological properties measured in shear, but that
there was some correlation with extensional flow properties. In a
later study [35], a similar result was found for both LDPE and
HDPE. The melt strength test described in Chapter 7 was found to
be the simplest method for detecting differences in drawability
between similar resins. Laun and Schuch [36] have reported a
similar observation.
Han and Kwack [37] compared three LDPE film resins and found
the maximum value of DDRXBUR, which they called the "blowability" of the resin, was best for resins having a narrow molecular
weight distribution and a low degree of long chain branching. In a
similar study of LLDPE film resins [38], the same authors found
that these resins were so easily drawn that rupture could not be
generated on the equipment available.

550

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

17.4 BUBBLE STABILITY

Instabilities in bubble behavior are usually the limiting factor in the


production of blown films. Since aerodynamic forces play a central
role in bubble stability, the limit often appears in the form of a
cooling rate above which bubble stability becomes a serious
problem.
Several types of instability have been observed in laboratory
studies involving the use of small scale film lines [39-42]. These
include the following:
1. Periodic thickness fluctuations, very similar in nature to the
draw resonance phenomenon in melt spinning [53], occur
when there is little inflation of the bubble and the DDR is
very high.
2. Periodic axisymmetric diameter fluctuations occur when the
product of BUR and DDR is high, but with a modest amount
of inflation.
3. Neck moves abruptly up and down at fixed time intervals.
4. Circular motion or wobbling of the entire bubble occurs at
high BUR and introduces a helical bulge in the final film.

Not all of these phenomena correspond to behavior observed in


production scale equipment.
There is also a high-frequency flutter or buffeting that results
from cooling air turbulence, but this does not seem to cause
production problems.
Cain and Denn [43] have carried out a mathematical analysis of
the stability of an isothermal film blowing process, assuming first
that the melt is Newtonian and then that it is viscoelastic and
described by the upper convected Maxwell model (see Section
3.5.2), which has a single relaxation time. Regimes of instability
were predicted for both constitutive equations when the internal
bubble pressure was fixed, but not when the bubble volume was
fixed. Due to the simplifications in the analysis and the use of
unrealistic constitutive equations, it is not possible to say with any
certainty whether these results are relevant to the instabilities
observed in commercial film blowing.
In its advanced stages, instability can lead to rupture of the
bubble and the interruption of production, and this can occur, for

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

551

example, in the case of very high-viscosity LDPE, because the


stresses are very high. However, it can pose a serious problem
without interfering with production by generating gauge variations
in the finished film. LDPE produces bubbles that are much more
stable than HDPE, and broadening the MWD of HDPE seems to
improve stability.
The difference between HDPE and LDPE is felt to be related to
the difference in their behavior in uniaxial extension experiments.
In particular, in start-up of steady simple extension, LDPE exhibits
a marked degree of extension thickening, i.e., its tensile stress
growth coefficient rises sharply above the linear viscoelastic curve
around a certain value of the Hencky strain. In a comparison of two
similar LDPE film resins [35], it was found that the resin with the
higher values of TJ +(t, i) and the lower maximum drawdown also
produced the more stable bubble. HDPE, on the other hand may
be mildly extension thickening but more likely extension thinning
(strain softening). It is hypothesized that the rapid build-up of stress
in the LDPE occurs also in the unbalanced biaxial stretching that
occurs in the film blowing process. If this is the case, the viscous
stress level in the melt will be higher compared to the aerodynamic
forces, and this will make the bubble less sensitive to the latter.
When LLDPE was introduced, it was intended as a replacement
for LDPE, and it was thus desirable to be able to process it using
equipment already in place for the latter. However, this could only
be achieved by reducing production rates, because the LLDPE
produced less stable bubbles than LDPE. The extensional flow
properties of LLPDE are more like those of HDPE than LDPE,
and it tends to be much less extension thickening than HDPE. In
order to classify film resins in this regard, Jones and Kurtz [32] have
proposed the use of an "extensional viscosity index" defined as
follows:

TJ;(S = 2.0; i = 1 S-l)


E.Y.I. == TJ; ( S = 0.2; i = 1 s- 1)

(17-8)

They define a "low-strain hardening" material as one in which


E.Y.I. is less than 4.5 and note that commercial LLDPEs have
values that are less than 3.5 in order to achieve high drawability.
For LDPE, on the other hand, the E.Y.1. can be as high as 7 or 8.

552

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

It has been found that one can achieve good bubble stability in
LLDPE at high production rates by the use of specially designed air
rings that take advantage of aerodynamics to control the shape of
the bubble. In particular, a dual-lip air ring has been found to be
advantageous [32].
Another approach to improving the bubble stability for LLDPE is
to use a blend containing LDPE. Huang and Campbell [27] found
that the addition of 10% LDPE changes the strain rate in the
molten portion of the bubble. However, the use of additional LDPE
had little further effect.
Ghijsels et al. [44] found a remarkable synergistic effect of
blending linear and branched polyethylenes having similar molecular weights and 710 values. Both the zero-shear viscosity and the
melt strength for blends had significantly higher values than for
either blend component. This synergistic effect was found to correlate with bubble stability as observed in actual film blowing [45].
Yet another approach to stabilizing LLDPE bubbles is to use a
"booster," which is a resin additive that alters melt flow behavior.
One possible type of booster is a peroxide crosslinking agent,
although care must be taken not to produce sufficient crosslinking
to interfere with the drawdown process or to produce gels.
Crosslinking would make LLDPE behave more like LDPE.
17.5 SHEET EXTRUSION

Another process for producing a continuous thin sheet of polymer


involves the use of a long slot die with a narrow gap. Such an
arrangement is the basis for slot casting (chill roll casting), embossing, and extrusion coating operations. In each case, the molten
extrudate, called a "web," is subjected to machine direction forces
because of the rotation of a drawdown roll that is part of each of
the above mentioned operations.
The rheology-related phenomena that can cause problems in
sheet extrusion are edge bead, neck-in, draw resonance and web
rupture. Extrudate swell occurs and plays some role in each of
these phenomena but is not, in itself, a problem.
An edge bead is a thickening of the web at its outer edges, which
can result in an edge that is as much as five times as thick as the
center of the web. This results in a nonuniform product and causes

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

553

problems in downstream nip flow and winding processes, and edge


bead is therefore usually trimmed off to become scrap or regrind.
Extrudate swell and surface tension contribute to the edge bead
problem. However, surface tension is a significant factor only in the
case of low viscosity coating resins, and it has been argued by
Debroth and Erwin [46] that the principal cause of edge bead is the
unique state of stress of the melt at the edges of the web.
While the melt away from the edge is subjected to a transverse
component of stress transmitted to it through the web from the
melt flowing through the outer portions of the die, the melt at the
edge has a free surface on which the normal stress is just that due
to ambient pressure. The melt in this region is subjected to a stress
field that is close to that of tensile extension, while the stress field
near the center of the web is more like that of planar extension.
The principal stretch ratio in both regions is approximately the
ratio of the roll velocity to the die exit velocity, VR/Vo . For planar
extension of an incompressible material, the thickness at the roll,
h R' is related to the stretch ratio as follows:

(17-9)
For simple extension, however, the diameter of a fluid element is
given by:

(17-10)
Thus, the edge bead thickness, dR' is related to the center web
thickness, h R' as follows:

(17-11)
Dobroth and Erwin [46] show data for a LDPE that agree with this
result.
Edge bead can be controlled to some extent by the use of special
die features such as inserts [47] and variation of the die gap opening

554

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

along its length [48]. However, some edge bead may be desirable to
prevent edge tear [49].
Neck-in is particularly troublesome in coating operations, because of the longer distance between the die and the substrate.
Neck-in reduces the width of the coating and is thus undesirable.
Any fluid element subjected only to a tensile stress in one direction
will shrink at an equal rate in all directions normal to this. In the
case of an extruded web, the only force resisting this tendency is
the lateral stress transmitted through the web from the melt at the
edges of the die. The neck-in process reflects the relative stress and
the relative resistance to deformation in the thickness and transverse directions. In general, it is found that a lower viscosity (lower
MW or higher temperature) produces a higher level of neck-in [50].
The melt strength test described in Section 6.3.2 has been widely
used to evaluate coating resins [51], and it is generally found that a
high melt strength correlates with a low degree of neck-in.
Draw resonance is a phenomenon that produces a periodic
variation in cross-section of an extrudate whenever it is drawn down
between two fixed points above some critical draw ratio, which is
the ratio of the velocity at the take-up roll divided by the die exit
velocity [52]. It causes nonuniform gauge and can lead to rupture
of the web. This is a hydrodynamic stability phenomenon that occurs in both melt spinning and sheet extrusion, even in the case of
Newtonian fluids [53,54]. The critical draw ratio for a filament
of Newtonian fluid is 20, while values observed in sheet extrusion of
polyethylene are often around 35 [55]. Draw resonance is much
more severe for linear than for branched polyethylene, and this was
an initial barrier to the penetration of LLDPE into several sheet
extrusion markets. One solution to this problem is to avoid the
sudden cooling of the web at the chill roll by use of a linear jet of
cooling air [55-57].
As the drawdown ratio is increased, the deformation rate and the
tensile stress in the web also increase, and the web will eventually
rupture. By generating regions with thinner than average thickness,
draw resonance will clearly exacerbate this problem. The maximum
drawdown ratio has been found to decrease with increasing "melt
strength" [50], a quantity that reflects the level of tensile stresses
generated in a stretching flow.

ROLE OF RHEOLOGY IN FILM BLOWING AND SHEET EXTRUSION

555

REFERENCES

1. E. L. Steward and A. W. Cline, "Barrier screw hikes quality of HMW-HDPE


blown film," Plastics Engineering, Sept. 1987, p. 45.
2. J. P. Christiano, SPE Tech. Papers 35:96 (1989); Plastics Eng., June 1989,
p.57.
3. S. J. Kurtz and L. S. Scarola, Plastics Engineering, June 1982, p. 45.
4. J. C. Miller, R. Wu and G. S. Cieloszyk, Plastics Engineering, Jan. 1986, p. 37.
5. B. Proctor, "Flow analysis in extrusion dies," SPE Journ. 28, Feb. 1972, p. 34.
6. J. Wortberg and K. P. Schmitz, Kunststojfe 72:198 (1982).
7. P. Saillard and J. F. Agassant, Polyrn. Proc. Eng. 2:37 (1984).
8. J. Vlcek, V. Kral and K. Kouba, Plast. Rubber Proc. App. 4:3099 (1984).
9. C. Rauwendaal, Polym. Eng. Sci. 27:186 (1987).
10. J. Vlcek, J. Vlachopoulos and J. Perdikoulias, Intern. Polym. Proc. II 3/4,174
(1988).
11. J. Perdikoulias, J. Vlcek and J. Vlachopoulos, Adv. Polym. Techno/. 7, no. 3,
333 (1987).
12. D. M. Kalyon, J. S. Yu and c.-c. Du, Polym. Proc. Eng. 5:179 (1987).
13. P. C. Gates, TAPPI J. 70 (no. 6):38 (1987).
14. F. N. Cogswell, Polymer Melt Rheology, John Wiley & Sons, New York, 1981,
p. 101.
15. S. J. Kurtz, in Advances in Rheology, vol. 3, p. 399, Edited by B. Mena et aI.,
UNAM, Mexico City, 1984 (Proc. IXth Intern. Congr. Rheo!.).
16. S. J. Kurtz, T. R. Blakeslee, III and S. S. Scarola, U.S. Patent 4,282,177 (1981).
17. "Heated die-lip system increases LLDPE film productivity," Modem Plastics,
Feb. 1987, p. 82.
18. A. Rudin, J. E. Blacklock, S. Nam and A. T. Worm, SPE Tech. Papers 32:1154
(1986).
19. A. J. Athey, R. C. Thann, R. D. Souffle and G. R. Chapman, SPE Tech.
Papers, 32:1149 (1986).
20. A. V. Ramamurthy, J. Rheol. 30:337 (1986).
21. A. V. Ramamurthy, U.S. Patent 4,552,712 (1985).
22. A. V. Ramamurthy, U.S. Patent 4,554,120 (1985).
23. A. V. Ramamurthy, U. S. Patent 4,522,776 (1985).
24. M. Rokudai, S. Mihara and T. Fujiki, J. Appl. Polym. Sci. 23:3289 (1979).
25. H. Miinstedt, Colloid Poly. Sci. 259:966 (1981).
26. R. Farber and J. Dealy, Polym. Eng. Sci. 14:435 (1974).
27. T. A. Huang and G. A. Campbell, Adv. Polym Techno/. 5 (3):181 (1985).
28. R. K. Gupta, A. B. Metzner and K. F. Wissbrun, Polym. Eng. Sci. 22:174
(1982).
29. B. Cao and G. A. Campbell, Intern. J. Polym. Proc. 4:114 (1989).
30. H. O. Corbett, U.S. Patent No. 3,167,814 (1965).
31. F. J. Herrington, U.S. Patent No. 4,118,453 (1978).
32. D. N. Jones and S. J. Kurtz, U.S. Patent 4,330,501 (1982).
33. T. Kanai and J. L. White, J. Polym. Eng. 5:135 (1985).

556

34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50:>
51.
52.
53.
54.
55.
56.
57.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

J. Meissner, Pure Appl. Chem. 42:553 (1975).


H. H. Winter, Pure Appl. Chem. 55:943 (1983).
H. M. Laun and H. Schuch, 1. Rheol. 33:119 (1989).
C. D. Han and T. H. Kwack, 1. Appl. Polym. Sci. 28:3399 (1983).
T. H. Kwack and C. D. Han, 1. Appl. Polym. Sci. 28:3419 (1983).
C. D. Han and J. Y. Park, 1. Appl. Polym. Sci. 19:3291 (1975).
C. D. Han and R. Shetty, IEC Fund. 16:49 (1977).
W. Minoshima and J. L. White, 1. Non-Newt. Fl. Mech. 19:275 (1986).
J. L. White and H. Yamane, Pure Appl. Chem. 59:193 (1987).
J. J. Cain and M. M. Denn, Polym. Eng. Sci. 28:1527 (1988).
A. Ghijsels, J. J. S. M. Ente and J. Raadsen, in Integration of Fundamental
Polymer Science and Technology-2, Ed. by P. J. Lemstra and L. A. Kleintjens,
Elsevier Applied Science, London and New York, 1988, p. 466.
A. Furumiya, Y. Akana, Y. Ushida, T. Masuda and A. Nakajima, Pure Appl.
Chem. 57:823 (1985).
T. Dobroth and L. Erwin, Polym. Eng. Sci. 26:62 (1986).
W. F. Allen, SPE Tech. Papers 33:211 (1987).
T. Dobroth and L. Erwin, SPE Tech. Papers 32:843 (1986).
R. Edwards, TAPPII. 70, no. 9:139 (1987).
E. J. Kaltenbacher, J. K. Lund and R. A. Mendelson, SPE loum., Nov. 1967,
p.55.
R. L. Ballman, Rheol. Acta 4:137 (1965).
A. Co, V. Iyengar and C. M. Lin, Xth Int. Congr. Rheol. 1:278 (1988).
C. J. S. Petrie and M. M. Denn, A.I.Ch.E.l. 22:209 (1976).
N. R. Anturkar and A. Co, 1. Non-Newt. Fl. Mech. 28:287 (1988).
P. J. Lucchesi, E. H. Roberts and S. J. Kurtz, Plastics Engineering, May 1985,
p.87.
P. J. Lucchesi, E. H. Roberts and S. J. Kurtz, U.S. Patent 4,486,377 (1984).
E. H. Roberts, P. J. Lucchesi and S. J. Kurtz, Adv. Polym. Technol. 6:65
(1986).

Chapter 18
On-Line Measurement of
Rheological Properties

18.1 INTRODUCTION

There are three types of application of on-line rheometers: process


monitoring, quality control, and automatic process control. The first
type arises primarily in the development of a new material or
process, when it is useful to be able to monitor the effects of
changes in formulation or operating parameters on product characteristics. In a quality control application, a slow and labor-intensive
laboratory test procedure is replaced by a direct indication of
product quality. The most sophisticated type of application is in
automatic process control, where the rheometer is used as a sensor
providing an input signal to the controller. A general review of
process rheometers and their applications has been published [1].
The rheological properties of an in-process material may be of
interest for its own sake, or as a measure of some other attribute
such as molecular weight. For some resins, such as "controlled
rheology" types or high-flow injection molding materials, the viscosity is the primary property that distinguishes different grades, so if
the resin manufacturing process could be controlled to produce a
preset viscosity, quality control would be automatic. More often, the
rheological property is of interest as a measure of some other
attribute. For example, the storage and loss moduli might serve as a
measure of the molecular weight distribution, or the viscosity might
be used as a measure of composition or the state of dispersion of a
compound or blend.
557

558

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

18.2 TYPES OF ON-LINE RHEOMETERS FOR MELTS

It is common practice to distinguish between "on-line" and "in-line"

instruments. An on-line rheometer makes use of a side stream and


a gear pump to carry a sample of melt to the actual sensor, whereas
an "in-line" rheometer is one that senses the properties of the melt
as it passes a particular point in the process. An "off-line" measurement is one made in a laboratory.
It is more difficult to design a rheometer for in-line use, but such
an instrument has the advantage of a short signal delay, and it
avoids the use of gear pumps. On the other hand, the sample
cannot be brought to a prescribed temperature prior to the measurement. Since fluctuations in temperature are a normal part of
any industrial process, this means that a temperature compensation
procedure must be used to reduce the measured property to a
standard temperature. If the material being processed obeys timetemperature superposition, this is quite straightforward, once the
shift factor is known as a function of temperature.
18.2.1 On-Line Capillary and Slit Rheometers for Melts

As is the case in the laboratory, the siraplest and most common


type of process rheometer is the capillary rheometer. A side stream
is taken from the main process line and forced through the capillary
by a gear pump. The capillary is thermostatted to control the melt
temperature. The flow rate is calculated from the speed of the gear
pump, and the pressure drop is measured directly by means of
pressure transducers. By cycling between several pump speeds, the
apparent viscosity at several shear rates can be determined.
If the rheometer exits to the atmosphere, only a single vented
pressure transducer is required. However, in this type of installation, all the material that enters the rheometer flows out as scrap at
a rate of about 0.5 to 1 kg/h. If this is to be avoided, a second gear
pump is required to return the capillary efflux back to the main
process stream. A second pressure transducer is then required, and
the speed of the second gear pump must be carefully controlled to
maintain a reasonable and constant level of pressure in the capillary.

ON-LINE MEASUREMENT OF RHEOLOGICAL PROPERTIES

559

Clearly it is not possible to carry out all the measurements necessary to determine the entrance corrections and the Rabinowitch
correction, and therefore on-line capillary rheometers are generally
used to indicate either the apparent viscosity or to simulate a melt
index measurement. The melt index test is simulated as follows.
The pressure drop is set at 298.2 kPa, which is the nominal value
for the melt index test, and a die having the prescribed L / D is
used. Then the motor driving the gear pump is regulated to maintain the pressure drop at the set level, and the resulting pump
speed is noted. The simulated melt index value is calculated as
follows:

(18-1)
where:

MI = melt index in g/10 min


V = pump displacement, cc/rev.
S = pump speed, RPM
K = empirical calibration factor to account for all the
differences between a melt indexer and the capillary
rheometer
p = melt density
Dm = standard die diameter for melt index test
Dc = capillary diameter

In practice, the product KVp is determined experimentally for the


particular resin of interest. The major manufacturers of capillary
on-line rheometers are Gottfert [2] and Seiscor.l
On-line rheometers based on slit flow have also been used. Flow
in a slit is explained in detail in Section 8.3, and the equations
necessary to calculate quantities of rheological significance are
presented there. The advantage of the slit over the capillary is that
it is possible to mount pressure transducers flush with the wall. This
has been of special interest because, as explained in Section 8.3.2, it
has been proposed that the first normal stress difference can be
determined from wall pressures measured in a slit. Both the hole
iAddresses of manufacturers are given in Appendix E.

560

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

pressure method and the exit pressure method are described there,
and commercial on-line sensors based on both these methods are
available? However, as explained in Chapter 8, questions remain
about the reliability of these two methods, and there have been no
published reports of their successful use for the on-line measurement of N 1 Another pressure-flow device that can, in principle, be
used to measure Nl is the curved slit rheometer developed by
Geiger [3]. As in the case of the previously mentioned slit rheometers, however, it is not clear whether this device can form the basis
of a useful industrial sensor.
The Rheometries Melt Flow Monitor is a commercial slit
rheometer that was developed not to measure Nl but to eliminate
certain disadvantages of other on-line instruments. The problems
addressed were:
1. Long signal delay due to long side stream
2. Waste stream produced by rheometer efflux
3. The need to mount major components near processing machinery.
This rheometer is equipped with two gear pumps, each driven
by its own stepper motor, with the speed of the downstream
motor regulated to provide a preset pressure at the exit of the
capillary. The entire unit is designed to be installed just above a
special adapter plate that can be placed at the end of an extruder
or incorporated into the barrel assembly so that melt is sampled
before it reaches the end of the extruder.
18.2.2 Rotational On-Line Rheometers for Melts

In a rotational on-line rheometer, a side stream from a process line


carries melt into the rheometer test chamber in which a rotating
element is mounted, and means are provided to measure the torque
on either the rotating or stationary element. The flow through the
rheometer can be continuous, although the through-flow will have
an effect on the measured torque, and more reliable results are
2The Seiscor/Han rheometer makes use of the exit pressure and is offered by Seiscor. The
Lodge Stressmeter is based on the hole pressure method and is offered by Bannatek.
Addresses of all manufacturers mentioned in this book are listed in Appendix E.

ON-LINE MEASUREMENT OF RHEOLOGICAL PROPERTIES

561

obtained by periodically stopping the flow to make a measurement.


Although the response of such an instrument is relatively slow, it
represents a vast improvement over periodic laboratory testing.
One challenge in the design of a rotational process rheometer is to
provide a method for measuring the torque without interference
from a static or dynamic seal. This problem is usually solved by use
of a torsion tube, which transmits torque while simultaneously
providing a static seal.
In the Rheoprocessor developed by Kepes [4, 5] the rotating
member is a bicone, and the speed or displacement arising from the
application of a constant or time-varying torque is monitored. A
seal is not necessary as the rheometer chamber drains to the
atmosphere by gravity and is thus not under pressure. A four-way
valve allows the periodic stopping of the flow for measurement or
the use of a calibration standard.
The Rheometrics On-Line Rheometer is a concentric cylinder
device designed for the measurement of linear viscoelastic properties. A gear pump provides a controlled flow of melt into the
rheometer. The outer cylinder is made to oscillate sinusoidally, and
the torque on the inner cylinder is measured by means of a torque
tube assembly that provides a perfect seal. The dynamic seal used
for the shaft driving the outer cylinder introduces no error, since
the torque is not measured on this shaft. Starita and Rohn [6] have
described this rheometer, and Zeichner and Macosko [7] have
discussed its application for monitoring the molecular weight distribution of polypropylene.
In another concentric cylinder process rheometer described by
Heinz [8] the concentric cylinder sensor is a self-contained unit that
is inserted into the melt stream, either in an on-line or an in-line
installation. The sensor consists of three concentric, thin-walled
cylinders. The second (middle) cylinder is driven in an oscillatory
mode, and a shear stress is generated on both sides due to the
presence of the two stationary cylinders. The torque on the drive
shaft is transmitted to an external displacement sensor by means of
a torque tube. Bellows are used to provide seals for the moving
parts.
Khachatryan et al. [9] have described a concentric cylinder
rheometer for use with fiber-forming resins. Two thin-walled concentric cylinders are suspended in a chamber through which the

562

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

polymer is made to flow. The inner cylinder is rotated by a motor,


and the torque transmitted to the outer cylinder is monitored. The
published report does not explain how the outer cylinder is suspended nor how the torque arm is sealed where it passes through
the wall of the chamber. Another concentric cylinder on-line melt
rheometer has been described by Menges et al. 3
18.2.3 In-Line Melt Rheometers

All of the on-line instruments described above require the use of a


side stream and at least one gear pump. In some installations,
where it is important to minimize the signal delay or when the use
of a gear pump is not practical, an in-line installation is advantageous. In Section 7.9.4, we mention a "shear stress transducer" that
has been developed for use in a sliding plate rheometer [10,11]. It
has been proposed [12] that such a device could be used as the basis
for an in-line rheometer. One possible configuration is shown in
Figure 18-1. A rotating drum (A) generates a nearly uniform
shearing deformation in the small gap (B), and the resulting shear
stress is monitored by the shear stress transducer (C). The motion
of the rotor also generates a drag flow that draws melt into the gap
to aid the process of sample renewal. A rheometer of this type has
been used to control reactive extrusion processes [13,14].
A single "flight" machined onto the drum could be used to
ensure a clean sweep of the shearing gap as often as desired to
ensure complete sample renewal and to avoid the build-up of a
degraded layer of material at the outer wall. Using a servo-motor to
drive the drum, any desired shear history can be generated, including steady shear, oscillatory shear and step strain. Thus, not only
can the viscosity be monitored but also the storage and loss moduli
and even nonlinear viscoelastic properties. By using a cam-like
drum having two or more diameters over various sections of its
perimeter, the viscosity at two or more shear rates could be measured once each revolution by simply driving the drum at a constant
rotational speed.
Another possible application of a shear stress transducer for
in-line measurements is the measurement of the shear stress on the
3 G.

Menges,

(1989).

w.

Michaeli, C. Schwenzer and L. Czybarra, Plastverarbeiter 40, No. 4:207

ON-LINE MEASUREMENT OF RHEOLOGICAL PROPERTIES

563

Figure 18-1. An in-line rotational rheometer that makes use of a shear stress transducer, as
described in Ref. 12. A = rotary drum driven by constant speed motor or servo-motor;
B = gap in which concentric-cylinder drag flow is generated; C = shear stress transducer
that senses the shear stress at the outer wall of the concentric-cylinder drag flow zone.

wall of an extruder barrel. Because the flow field is neither simple


nor under complete control, this would not provide information
about any well-defined rheological property. However, if a special
"rheometric module" were inserted into the screw, and if the screw
speed were accurately monitored, it would be possible to make a
measurement of the viscosity.
Revesz [15] has described a device for measuring the shear stress
exerted by the melt on the tip of an extruder screw. He suggests a
method for estimating the viscosity of the melt as it enters the die
[16], although the principal purpose of the device is to provide
information about melt flow rate.
18.3 SPECIFIC APPLICATIONS OF PROCESS RHEOMETERS

Some types of processes in which process rheometers have potential


applications are:
1.
2.
3.
4.
5.

Polymerization
Extrusion and melt forming
Compounding
Blending
Reactive extrusion.

564

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

The advantages to be gained by the use of such sensors include


improvements in product quality, reduction of waste, and increases
in production rates.
Possible manipulated variables in process control applications
are:
1. Feed rates of reactants or additives
2. Process temperature
3. Extruder speed
4. Geometry of die or other mechanical components.

Feed rate can be controlled by a liquid pump or by a loss-in-weight


solids feeder, and this is the obvious choice of manipulated variable
in reactive extrusion operations. For example, it may be desirable to
accomplish a controlled degree of degradation or of crosslinking to
produce an extrudate having a specified viscosity or elasticity.
Temperature is not truly a possible manipulated variable, as it is
actually changed by manipulation of either the thermal or mechanical energy input to the process. Processing machinery usually has a
large thermal inertia, and this leads to a slow response to a change
in the temperature set point if thermal energy input is used as the
manipulated variable.' Changing screw speed alters the rate of
mechanical energy input, and this is a much faster way of changing
the temperature. However, this also changes the flow rate, pressure,
and strain history of the melt. Changing machine geometry, for
example by means of a hydraulically variable die gap, can change
the rheological properties of a shear sensitive polymer such as
LDPE.
An obvious area of application is in the polymerization process
itself, although there have been no published reports of such uses.
If the product of a polymerization process is itself a melt, then a
process rheometer can be used to monitor this product directly.
However, several modern polymerization processes produce a powder rather than a melt, and in order to use rheology as a tool for
process control it is necessary to provide a mechanism for automatically sampling this stream and converting powder to melt. The
Rheometries ROR-QC addresses this problem by use of pneumatic
transport [6].
Starita and Macosko [17] have discussed some ways in which
variations in resin characteristics can cause problems in melt form-

ON-LINE MEASUREMENT OF RHEOLOGICAL PROPERTIES

565

ing operations such as blow molding. If a process rheometer could


provide continuous information regarding the nature of the raw
material, operating conditions could be altered to compensate for
such changes.
Variations in viscosity can cause serious difficulties in an injection
molding operation, especially when large parts are being produced,
and it would be useful to be able to monitor melt consistency on a
continuous basis. At present, nozzle pressure is sometimes used as
a rough indication of viscosity [18].
The use of compounding operations to produce sophisticated
multicomponent materials is growing rapidly, and there is a widely
recognized need for methods to ensure that the product has the
correct formulation and is properly mixed. Process rheometers
show promise in this area, specifically to monitor the state of
dispersion of a masterbatch, a compound or a blend [3, 6].
Considerable attention has been focussed on the use of process
rheometers for the closed-loop control of reactive extrusion processes. If a rheological property is sensitive to the extent of reaction, then a process rheometer can be used to provide a signal for
the control of the feed rate of one of the reactants. Several reports
have been published on the use of rheometers to control the
peroxide-initiated degradation of polypropylene [19]. This is a process that is used to produce a range of lower molecular weight
resin grades starting from a single high-molecular weight grade
using an extruder. Both rotational [7,20] and capillary [21] on-line
rheometers have been evaluated for this application. In addition,
Pabedinskas et al. [22] used the die pressure drop as a measure of
the amount of degradation.
REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.

J. M. Dealy, Chern. Eng., Oct. 1984, p. 62.


A. G6ttfert, KunststoJfe 76:1200 (1986).
H.-G. Fritz, KunststoJfe 75:785 (1985).
A. Kepes, Proc. 8th Int. Congr. Rheol. 2:185, Plenum Press, New York, 1980.
A. Kepes, U.S. Patent 4,334,424.
J. Starita and C. L. Rohn, Plastics Compounding, Marchi April 1987, p. 46.
G. R. Zeichner and C. W. Macosko, SPE Tech. Papers 28:79 (1982).
W. Heinz, Proc. IXth Intern. Congo Rhea!. 4:85 (1984).
G. M. Khachatryan, K. D. D'yakov, A. A. Strel'tsov and K. N. Sosulin, Fibre
Chern. 15:228 (1983), Translation from the Russian of Khimicheskie Volokna,
No.3, p. 48, May-June, 1983.

566

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

J. M. Dealy, U.S. Patent No. 4,463,928 (1984).


A. J. Giacornin, T. Sarnurkas and J. M. Dealy, Polym. Eng. Sci. 29:499 (1989).
J. M. Dealy, U.S. Patent No. 4,571, 989 (1986).
T. O. Broadhead, Doctoral Dissertation, Chern. Eng., McGill Univ., Montreal,
1991.
B. Nelson, Doctoral Dissertation, Chern. Eng., McGill Univ., Montreal, 1991.
H. Revesz, Kunststojfe 64:35 (1974).
L. Halasz, I. Mondvai and H. Revesz, Rheol. Acta 22:313 (1983).
J. M. Starita and C. W. Macosko, SPE Tech. Papers 29:522 (1983).
A. R. Agrawal, I. O. Pandelidis and M. Pecht, Polym. Eng. Sci. 27:1345 (1987).
A. Dreiblatt, H. Herrmann and H.-J. Nettelnbreker, Plastics Engineering, Oct.
1987, p. 31.
H.-G. Fritz and B. Stoehrer, Intern. Polym. Proc. 1:31 (1986).
J. Curry, S. Jackson, B. Stoehrer and A. van der Veen, Chem. Eng., Nov. 1988,
p.43.
A. Pabedinskas, W. R. Cluett and S. T. Balke, Polym. Eng. Sci. 29:993 (1989).

Chapter 19
Industrial Use of Rheometers
19.1 FACTORS AFFECTING TEST AND INSTRUMENT SELECTION

Even with unlimited resources of equipment and personnel it would


be necessary to make choices as to which of the infinite number of
conceivable rheological properties to measure and which instruments to use to measure them. In the real world limitations of
resources, and especially of time, make these choices even more
necessary.
Of course there is no unique answer to the question "Which tests
should I perform on this material and with which rheometer(s)?"
One reason for this is the variation in the availability of human and
material resources. Different answers would be appropriate, for
instance, if circumstances required the use of existing instrumentation, as opposed to a situation that permitted the acquisition of new
equipment. In some cases the process of interest may be so complex
that appropriate test methods have not yet been developed. Rather
than giving specific answers, we provide here some guidelines that
the reader should find useful in formulating his or her own answer
to the question.
It is appropriate to introduce this discussion with a quote from a
discussion of test methods by Mooney, a pioneer in the application
of rheology to polymer processing [1]:
"Methods of testing materials may be classified as being either purely
'practical,' purely 'scientific,' or hybrid. The ideal 'practical' test is a
service test, or test by trial in the use for which the material is intended.
Service tests are usually expensive, slow, and so inaccurate that many
are required to give a reliable average result. Laboratory tests must
therefore often be used. The best laboratory 'practical' test is the one
567

568

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

which imitates most faithfully some actual seIVice condition while


giving, at the same time, quick and accurate results.
"A 'scientific' test is one which measures some specific physical
property of the material. In order for the test to do this, it is essential
that the measurement shall be unaffected by extraneous physical properties and shall be expressible in absolute units, independent of the
design and dimensions of the testing equipment.
"A 'hybrid' test is one that stands somewhere between the 'practical'
and the 'scientific.' It does not represent faithfully any particular seIVice
condition; and it is not designed so as to permit the calculation of any
basic physical property. Most such tests in common use have been
designed with half an eye on seIVice conditions, and an eye and a half
on speed and accuracy of the laboratory test. If the measurements have
any scientific value, it is largely a result of chance, or oversight.
"There are many seIVice conditions which are very complex in
geometrical form or other aspects. Hence, it often is impossible to
analyze these seIVice conditions theoretically, or to predict just how and
to what extent the various physical properties affect the results of a
seIVice test. Complicated problems may sometimes be solved by a
combination of incomplete theory, dimensional analysis, and physical
testing. However, until such a program has been carried out, in difficult
problems the engineer or technologist can rely only upon practical tests
for a final decision or rating. He should make his practical laboratory
test resemble as closely as feasible the seIVice condition he is interested
in; and he owes no apology to the scientific world for doing so .... "

We can only add that Mooney's words, written in 1937, are still
applicable today.

19.1.1 Purposes of Rheological Testing

In selecting a test method, the first question to be addressed is:


what is the purpose of the test and what action will be taken on the
basis of the results? The most common applications of rheological
data are in the following areas:
1. Laboratory Characterization. Rheological tests are useful for
comparing the properties of materials, for determining their
structure, and for estimating their potential usefulness.

INDUSTRIAL USE OF RHEOMETERS

569

2. Processing. Typical applications are in the selection of resins


for a given process, the optimization of formulations, and the
diagnosis and solution of processing problems. Rheological
data are also required for process simulations.
3. Quality Control. Both resin manufacturers and users need to
ensure that properties critical for a processing and/or end use
are maintained within acceptable limits.
Different instruments and test protocols are generally best suited
for each type of application, and these are therefore discussed
separately in Sections 19.2 through 19.4.
19.1.2 Material Limitations on Test Selection

The material under study is a factor to be considered in selecting


the rheological property to be measured and the instrument with
which to conduct the test. The material may have particular properties that dictate choices not attractive solely on rheological grounds.
For example, if the material is extraordinarily sensitive to oxidation,
the rheometer must permit loading and testing in the absence of
air. Or, if the polymer has limited thermal stability in the melt, the
rheometer must be able to equilibrate temperature and make
measurements rapidly to minimize degradation.
The rheological nature of the material may itself impose constraints on the selection. Consider, for example, measurement of
the viscosity of a high melting polymer with a very low viscosity.
Rotational instruments are of limited use, because the low viscosity
requires large diameter fixtures in order to achieve sufficient sensitivity, and it is difficult with many instruments to achieve good
temperature control and uniformity with large fixtures. Capillary
viscometry with a piston driven instrument and load cell force
measurement also offers difficulties. The piston clearance is critical
for the measurement of low viscosity melts. Too small a piston
permits backflow, but if the piston is even slightly too large, friction
becomes comparable to the force required to extrude the melt. A
piston-driven instrument that incorporates a pressure transducer,
and is therefore not affected by piston friction, is one possible
solution. Another is to use a gas-pressure-driven rheometer with a

570

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

loose follower plug on top of the melt to prevent channeling of the


driving gas.
As another example, consider the problems encountered when a
dense filler is added to this low viscosity melt. Now the problem is
not a low signal level, because the filler raises the viscosity, but the
problem of maintaining a uniform concentration of filler in the test
sample. Settling by gravity and particle migration in the flow field
can both cause segregation of the filler. Use of a torque rheometer,
in which the test fluid is constantly agitated and mixed, may be
expedient in this situation. However, the complex flow field in such
a device makes it impossible to interpret the results in terms of well
defined rheological properties [2, p. 168].
Materials whose rheology is sensitive to deformation history over
time scales comparable to or longer than the normal time for a
measurement present another challenge. Inevitably the very act of
preparing the test specimen and loading it into the rheometer
imposes a history. It is especially important in this case to consider
the purpose of the test and how the results are to be used.
Standardization of the sample preparation procedure, e.g., by premolding test shapes under controlled conditions and loading these
with minimum additional deformation, may be adequate in some
cases.
In other applications, the simulation of the deformation history
of a process may be useful. For example, for a material whose
rheology is affected strongly by extrusion, the use of an extruder-fed
capillary rheometer might be advisable. A refinement of this technique is to pass the material through an adaptor in which a
controlled amount of shear, at a predetermined shear rate, is
applied uniformly to the melt before it enters the capillary rheometer [3].
Another tactic is to impose a sufficiently strong shear history to
erase all memory of previous deformation. It is useful for this
purpose to use a rotational rheometer that can be operated in both
shear and oscillatory modes. The steady shear is used to create an
initial history, and the small amplitude oscillatory shear can then be
used to probe the melt with minimal further perturbation of its
state. Alternatively, a high frequency and/or high amplitude oscillatory shear can be used to establish a standard state [4].

INDUSTRIAL USE OF RHEOMETERS

571

Finally, a prosaic but crucial material limitation that affects the


choice of test and instrument is that of available sample size. This
limitation is encountered most often when characterizing a newly
developed material. In general, a drag flow instrument (rotational
or sliding surface), is likely to be the instrument of choice. A sample
size as small as 0.5 g is often suitable. If that is not acceptable, or if
even this small sample size is excessive, it may be necessary to
devise an apparatus expressly for the problem at hand. An example
is a capillary viscometer that requires only 0.015 g of material that
was designed to measure pellet-to-pellet variations of melt viscosity
[5]. This can be compared to the at least 3 g of sample required for
capillary rheometers.
19.1.3 Instruments

The principles of operation, range of measurement capabilities, and


sources of error for many melt rheometers have been reviewed by
Dealy [2]. He indicates that a valuable tool for elasticity measurement is a creepmeter, i.e., an instrument in which the applied stress
is controlled and the resulting deformation measured. Such instruments are now commercially available (see Chapter 7 and Appendix
E). However, simple and inexpensive testers for melt elasticity and
instruments to measure elasticity at high shear rates are still
lacking.
A trend that was already underway at the time of Dealy's review
and that has now become almost universal is the use of computers
for control of operation and data acquisition in all types of rheometers. This offers the obvious advantage of decreasing the time
required to make a series of measurements and analyze results. It
also permits operation by technicians with less skill than that
required for manual operation.
A disadvantage relative to the slower pace of manually controlled
operation is that there is less opportunity to make direct observations during the measurement. Is an extrudate smooth and regular,
or is it bubbly, suggesting either poor packing or gas evolution? Has
melt fracture occurred? In torsional flow rheometry, what is the
condition of the exposed edge of the melt? Is the stress waveform
from a dynamic experiment sinusoidal? Even if it is not, the data

572

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

acquisition system will compute G' and G" as though the signal
were sinusoidal. If not, does the test material have a yield stress, or
is the non-sinusoidal waveform merely a result of the test fixture
rubbing against the oven? Has a pressure or torque reading really
levelled off, or is it still changing but at a sufficiently low rate to be
considered acceptable by the data acquisition system? It should be
remembered that the level of acceptability of a signal is usually set
by pre-assignment of a parameter of the data acquisition program.
An inappropriate setting can lead to the obtaining of artifacts
rather than of meaningful data.
Another disadvantage of automated operation is that the computer will process the signal obtained without regard for its physical
validity. Take for example an oscillatory shear fr~quency sweep
programmed to operate at a constant strain amplitude. At low
frequencies the torque may be below the lower limit of linearity of
the transducer, and the signal-to-noise ratio too low to be meaningful. Analogously, in a programmed shear rate sequence with a
capillary rheometer the pressure may be too low to be measurable
at the lowest shear rates programmed. Nevertheless, in both cases
the data acquisition program will happily compute and print out the
resulting meaningless numbers and call them viscosities. At high
frequencies or shear rates, programmed settings can cause erroneous readings because of shear heating or compressibility effects.
More sophisticated software would be capable of evaluating conditions that give rise to spurious data and discard them. Better still,
the software should be capable of adjusting the rheometer operation to avoid these conditions. With some programs it is possible to
use the instrument computer to calculate critical parameters such
as the adiabatic heating rate during the test. These can then be
printed out with the data and serve as warnings of possible artifacts.
An important consideration in selecting an instrument for a given
purpose is the experimental range over which data are required. In
general, an instrument with a given geometry and force measurement device can cover only about three decades of shear rate with
adequate precision. By geometry we mean, for example, for a
capillary rheometer a combination of barrel diameter and capillary
length and diameter, or, for a rotational rheometer a diameter and
gap. If data are required over a wide range of conditions, an
important factor is the ease of changing fixtures and/or transducers

INDUSTRIAL USE OF RHEOMETERS

573

to extend the range. Alternatively, the use of entirely different


instruments for different conditions may be indicated. The least
satisfactory situation is to be forced to rely on data from both
extremes of the useful range of an instrument.
Another important criterion for selection of an instrument is its
ease of operation and reliability, relative to the degree of training
of the operators and the environment in which the instrument will
be used. Frequency of need for calibration or of standardization is
part of this consideration. These questions are discussed at further
length in Section 19.4 on Quality Control Tests, where they are
particularly important.
19.2 SCREENING AND CHARACTERIZATION
19.2.1 Advantages and Disadvantages of Rheological Tests

For a number of reasons rheological measurements are often the


method of choice for the characterization of polymers produced
from new monomers, or of polymers made from known monomers
by new processes. In addition, because so many polymers are
processed as melts into their final form, some knowledge of rheological behavior is necessary to assess processability.
The strong dependence of rheological properties on structure,
and the relative ease of making rheological measurements, compared to the absolute determination of molecular weight, are additional reasons for using rheological measurements as part of a
screening process. We recall, for example, the 3.4-power dependence of '110 on Mw' Two samples of a polymer that differ in
molecular weight by 10%, which is near the limit of discrimination
of routine molecular weight methods, have a readily detectable
difference in viscosity of nearly 40%. The sensitivity of the compliance to small amounts of high molecular weight components is even
more pronounced. Also, rheological measurements are applicable
to polymers that are either insoluble or soluble only in solvents that
are difficult or hazardous to handle, making the classical dilute
solution characterization techniques inapplicable. Wu's [6] estimation of the molecular weight distribution of polytetrafiuoroethylene
by inference from rheological data illustrates this point. Tuminello

574

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

[7] has also described a procedure for estimating molecular weight


and MWD from such data and gives references to earlier methods.
A disadvantage of rheological measurements for the characterization of MWD is that they do not provide absolute values, but
neither then do the most popular dilute solution methods, intrinsic
viscosity (see Appendix C) and gel permeation chromatography,
which are also relative methods. If absolute values are needed,
these techniques also require calibration against an absolute standard. All of these measurements are most useful, therefore, for
comparative purposes, for example, for determining the relative
effect of a change in the polymerization process or monomer
composition.
19.2.2 Other Information Useful for Screening

It cannot be overemphasized that the rheologist's opportunity to


obtain useful information on a new material is greatly enhanced by
knowledge of its chemistry and of other physical properties. For
example, it is necessary to know the glass transition temperature
(see Appendix D) of an amorphous polymer, or the melting point
(Tm) of a crystalline one, to set the lowest temperature for preparing a sample for viscosity measurement.
The transition temperatures are usually measured easily and
rapidly by Differential Scanning Calorimetry (DSC). The determination is preferably made on a second scan, after the sample has
been heated and cooled. This precaution is essential for a polymer
such as polycarbonate that, although crystallizable, does not crystallize by cooling from the melt. Sample preparation and measurement can be done, therefore, down to temperatures near Tg
Exposure to certain solvents, however, will cause crystallization to
occur. A sample that has been so exposed must be heated to a
temperature above Tm for a valid measurement to be made.
Knowledge of the transition temperatures is also useful for
avoiding the occurrence of pressure-induced solidification during a
measurement, as both Tg and Tm are increased by pressure. The
variation of Tg with pressure is discussed in Section 10.5.1 where
it is estimated to be about lSC per 1000 psi. The magnitude of
the pressure dependence of Tm can be estimated by use of the

INDUSTRIAL USE OF RHEOMETERS

575

Clapeyron Equation:
(19-1)

where .lV, .lS, and .lH are the changes of volume, entropy, and
enthalpy, respectively, of fusion. For example, for a polymer with a
Tm of 500 K, a .lV of 0.1 cm 3 /g, and a .lH of 20 caljg, a pressure
of 1000 psi would raise the melting point by 5 K. Inadvertent
solidification of the melt will produce erroneous data and may
cause breakage of a piston or other damage to the instrument. The
likelihood of this occurring is particularly high when using a velocity
controlled capillary rheometer with the melt near a transition
temperature.
When measuring the properties of mixtures or of block copolymers it is important to know the phase diagram. Multiple phase
boundaries may be present. A given composition may therefore be
immiscible below one temperature and above another, but miscible
at intermediate temperatures. Furthermore, the kinetics of the
phase transformations, whether mixing or separating, are often slow
enough so that the mixture can be quenched to a metastable state.
The phase structure of the system, and thus the rheological behavior, will depend on the thermal history in what may be a very
complicated fashion. Confusion can arise if this possibility is
ignored.
Another type of physico-chemical data that may be important is
information on the kinetics and equilibrium of sorption of low
molecular weight species that the polymer has or may come into
contact with. These include solvents that can decrease Tg or cause
crystallization. Also of concern are reagents, such as water, that can
hydrolyze condensation polymers, causing random chain scission.
Sorption data are also useful for designing drying studies, and for
ensuring that proper precautions are taken against recontamination
after drying.
Crystallization rate data may be necessary to set drying conditions for polymers that crystallize slowly, such as polyethylene
terephthalate (PET), and that may therefore be received as amor-

576

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

phous material. If such a material is dried near Tg the granules of


polymer may sinter, making it impossible to load it into a rheometer. An alternative is to crystallize the polymer at a temperature at
which crystallization is rapid enough to occur before sintering. It is
necessary to verify that the moisture in the sample has not caused
hydrolytic degradation during the crystallization step. In some cases
it may be necessary to devise a tedious multistage drying process to
avoid all the possible pitfalls.
Understanding the polymerization reactions and those that can
occur subsequently is also helpful to avoid errors in measurement
and in the interpretation of data. As already discussed, condensation polymers are generally prepared by reversible reactions. Care
must be taken to preserve the molecular weight of the material as
received. This requires maintaining contact with the concentration
of condensation product, such as water, that is in equilibrium with
that molecular weight under the measurement conditions.
Other more subtle equilibrium considerations can be important
when one is dealing with mixtures or with copolymers. Interchange
reactions can occur to produce new species. For example, two
otherwise immiscible homopolymers can react to form a copolymer,
resulting in a single-phase material [8]. On the other hand, the
distribution of comonomer units that was in equilibrium under
polymerization conditions may shift to a different equilibrium when
the polymer is heated. Phase separation and a resulting gross
change in rheology can result [9].
The possibility of long chain branch formation during polymerization or subsequent handling is important because of the profound
influence of branching on rheological behavior. Free radical polymerization provides the obvious branching mechanism for chain
transfer to the polymer. However, other polymerization processes
can offer less obvious routes. The addition of a growing chain to the
terminal vinyl group of linear polyethylene in the Phillips process is
an example. The chemical structure of the polymer can also give
hints as to the possibility of branch formation. A polymer with
readily oxidized functional groups can be branched by limited
exposure to oxygen at elevated temperatures. The possible presence
of multifunctional compounds in the reagent stream is another
source of branches.

INDUSTRIAL USE OF RHEOMETERS

577

Solid materials can also react with the polymer. These might be
catalyst residues or an additive, such as a stabilizer or nucleating
agent [10]. Ionization of otherwise neutral end groups is also known
to affect melt rheology [11].
Thermal stability is another property of the polymer about which
information is vital. Reactions that cause chain scission or extension
or branching have severe effects. These are perhaps best detected
by rheological means, as discussed in the next section. Other
reactions can also occur. Thermal decomposition may result in the
evolution of small molecules, without significantly changing the
molecular weight of the polymer.
Bubbles of volatiles can change the viscosity directly, and they
can also have an indirect effect on the measurement. For example,
they can force a melt out of the gap of a torsional flow instrument.
In an extreme case the pressure of the gas evolved has been
observed to lift the piston and weight of a melt indexer out of the
rheometer barrel! Thermogravimetric analysis (TGA) can be used
to investigate the likelihood of gas evolution. Temperature scanning
is the usual mode of running TGA, but an expanded scale isothermal measurement at the temperature of, and over the time scale of,
the rheological test may be more meaningful.
To recapitulate, the better understood the chemistry of the
system under study, the more likely it is that reproducible measurements characteristic of the system can be made and that these can
be interpreted meaningfully.
19.2.3 Stability

As indicated in the preceding section, changes in melt rheology are


sensitive measures of polymer stability. Stability data can be used:
1. To determine if the rheology of the test material has changed

during the course of the measurement and, if it has, to correct


flow curves for the change
2. To ascertain the effects of reagents introduced deliberately or
inadvertently
3. To determine allowable limits on processing conditions such
as temperature, residence time, and shear rate.

578

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

19.2.3. 1 Stability Measurement

Stability can be estimated with most types of rheometers. Rotational instruments have an advantage in that the test material
remains in the rheometer for the desired length of time. A dynamic
measurement is preferred to steady shear, because the sample is
less likely to be disrupted by edge fracture or other flow irregularities. The test frequency and strain amplitude used depend on the
material. The strain amplitude is usually kept below 50% to avoid a
nonlinear response. The frequency must be high enough for the
stress response to be above the minimum sensitivity level of the
transducer. It should also be sufficiently high so that the rheology
does not change appreciably during the time required for a measurement. On the other hand, it is advantageous to operate in the
low-frequency region of the complex viscosity curve; the sensitivity
of rheological parameters to structure variation is more sensitive
there than in the power-law region. If the material properties
change drastically during the course of the measurement, it may be
necessary to change the test conditions to maintain the instrument
responses within desirable limits.
Capillary rheometers can also be used for stability studies. They
have an advantage in some cases in that they are easier to load
without appreciable contact of the sample with air [12], In a
capillary instrument the sample used for each measurement is
extruded and measured only once. The effect of residence time in
the rheometer reservoir is determined by intermittent or steady
extrusion, at constant shear rate or constant shear stress, depending
on the rheometer design. Precautions regarding test conditions
similar to those discussed above are necessary. If shear degradation
is the instability of interest, it is possible to collect and reextrude
the test sample repeatedly. However, because in a Poiseuille flow
the shear rate varies with radius, it is preferable for this purpose to
use a drag flow rheometer in which all of the material experiences
the same amount of deformation [3].
Torque rheometers, such as those from Brabender or Haake, (see
Appendix E) can be useful in some cases. They are not the
instruments of choice for initial screening studies, because they
require fairly large amounts of material and because the torque
response is not a well defined rheological property. On the other

INDUSTRIAL USE OF RHEOMETERS

579

hand, they offer the possibility of introducing and mixing into the
melt a reactant, either a stabilizer or a degradant, and observing the
subsequent response of the material with time. They can thus be
useful in screening potential stabilizers and in preparing stabilized
material for testing in conventional rheometers.
So far we have made the tacit assumption that instability can be
observed by following the change of a rheological property with
time, but there are times when this is not so. When a structural
change occurs sufficiently rapidly, it may take place during the
sample loading and temperature equilibration time, before measurements can be started, and may therefore not be observed as a
further change. The possibility of such an occurrence should be
examined by means of some independent measurement of structure
made before and after the rheological test.
An example of this sort of problem is the hydrolytic chain scission
of condensation polymers, especially polyesters and polycarbonates.
The need to dry these materials before melt processing is well
recognized; manufacturers of these resins routinely stress this in
their product bulletins. What is often not appreciated is the difficulty of maintaining a low level of moisture to prevent degradation.
To illustrate this point we show in Figure 19-1 a calculated curve
for the relative change of melt viscosity of PET as a function of
time of exposure to air at 25C with a relative humidity of 50%. The
assumptions made in this calculation were that the initially dry PET
is in the form of spherical pellets having a diameter of 0.125 inch
(0.32 cm); that the polymer has an inherent viscosity of 0.65 dl/g,
corresponding to an Mn of 20,000 g/mol; that each water molecule
causes random scission of a chain, preserving the initial "most
probable" MWD, and that the viscosity follows the 3.4-power
relation with Mw' The amount of water absorbed for each time was
calculated from the data of labarin and Lofgren [13]. After five
minutes of exposure, a time scale that is on the order of that
required to load many rheometers, the melt viscosity had decreased
to 0.7 times its initial value. An experimental observation of this
effect has been reported by Wissbrun and Zahorchak [12].
We note that the wide spread in the results from the IUPAC
working party study of the rheology of PET [14] is probably due to
differences in the drying and handling techniques used in the
various participating laboratories.

580

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

1.0F======-------------,
0.8

S 0.6

-;::: 0.4

0.2

______________
1.0

______________
10

____________
10 2

10 3

TIME (s)

Figure 19-1. Calculated relative change of melt viscosity of amorphous PET versus time of
exposure to 50% relative humidity air at 23C. Assumptions: 1/8 in. diameter spheres, initial
Mn = 20,000 g/mol, random scission of initially most probable distribution.

19.2.3.2 Use of Stability Data

If the polymer is found to be unstable in the initial screening, an

effort should be made to determine and eliminate the cause of the


instability. As discussed above, knowledge of the chemistry of the
system is invaluable here.
Sometimes a practical means to stabilize the polymer is not
available. Or it may not be deemed worthwhile to embark on a
stabilization program until an initial evaluation of properties has
been made. If the rate of viscosity change is not excessive it may be
possible to determine a flow curve with acceptable accuracy by
correcting for the change. A technique for doing so is to make an
oscillatory shear frequency scan, beginning at a frequency low
enough to give a reasonable estimate of the zero shear viscosity.
During the scan the measurement at this initial frequency is repeated periodically to give an estimate of how 110 changes with
time. The times at which the higher frequency points are measured
are then recorded.

INDUSTRIAL USE OF RHEOMETERS

581

The flow curve from the raw data is erroneous, because the
points measured at various times are affected differently by the
instability. An approximate correction can be made if one assumes
that the shape of the flow curve remains unchanged during the
measurement. Only its position on the viscosity and shear rate axes
is shifted, and the amount of the shift is determined from the
change of 'YJo with time. Explicitly, it is assumed that at any instant
the flow curve is obtained from a master curve equation:
(19-2)
The uncorrected data are corrected for the change in structure
that is indicated by the observed change of 'YJo by means of the
operations indicated below:

, 'YJo(O)
'YJ (t) 'YJo{t)

We =

w(t)-(-)

'YJ~

'YJo( t)

'YJo 0

(19-3)

(19-4)

where 'YJ'(t) is the uncorrected dynamic viscosity measured at time t


at the frequency wet); 'YJo(t) is the zero shear rate viscosity interpolated to time t, and 'YJoCO) is 'YJo(t) extrapolated back to zero time.
The process is illustrated schematically in Figure 19-2.
The assumption of an invariant master curve is likely to be valid
if the polymer initially has a fairly narrow MWD and if the reaction
it undergoes is random chain scission or chain extension. The
extrapolation of 'YJo(t) to zero time causes some uncertainty if the
viscosity is changing rapidly because of the ambiguity of the zero of
time during the temperature equilibration period.
The determination of the stability of a rheological property is an
important part of the screening process. The result may be a
deciding factor in establishing the suitability of a new polymer for
an application. At least it will indicate what precautions need be
taken during polymerization and work-up. In conjunction with
chemical techniques it can provide clues to the mechanism of the
instability and suggest methods for its inhibition.

582

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

(a)
'10(0)

,C:==:::====:;:=====:::::::J

t,

TIME

(b)

'l~(t,)

log(w)

Figure 19-2. Sketch to illustrate method for correcting flow curve for degradation that occurs
during measurement, based on Equations 19-2 to 19-4: (a) Logarithmic plot of 7]o(t), where a
and b are the deviations from loge 7]0) interpolated to times t I and t 2 respectively; (b)
Construction of corrected curve of loge 7]') versus log(w). Values measured at times t I and 12
are shifted both horizontally and vertically by a and b respectively.

19.2.4 Temperature and Frequency Dependence


19.2.4.1 Measurement Tactics

Having addressed the question of stability, we turn to the determination of the flow curve and of its temperature dependence. Again,
various types of rheometers can be used, but dynamic measurement

INDUSTRIAL USE OF RHEOMETERS

583

with a drag flow instrument is preferred for screening. A measurement can be made rapidly and requires only a small sample. The
parallel disk geometry is especially suitable for exploring a range of
temperatures. The differential thermal expansion that occurs with a
temperature change has only a small effect on a 1 or 2 mm gap, and
it can usually be neglected for screening purposes. If desired, a
correction can be made on the basis of a calibration of the effect of
temperature on the spacing. The cone-plate geometry is less tolerant of temperature variation because of the very small gap near the
vertex of the cone; thermal expansion can bring the cone and plate
into physical contact.
With a drag flow instrument it is possible to measure the elasticity as well as the viscosity of the melt. If such an instrument is not
available, a capillary rheometer must be used to determine the flow
curve. In that case the extrudate swell should always be recorded as
well. This requires little extra effort and gives a qualitative indication of melt elasticity. The end correction can also be measured.
This will insure that the die used has an L / D ratio large enough to
give an accurate value of the viscosity. In addition, the end correction is sensitive to the structure of the melt and may itself be useful
in the screening process.
In order to minimize stability problems it is best to start a
sequence of tests at the lowest practical temperature. After running
a frequency scan the temperature is raised at convenient intervals
and the procedure repeated. If the melt viscosity is very low it is
advisable to limit the frequency of measurement in order to avoid
centrifuging the sample out of the gap. It is a good idea to repeat
the measurement at the initial temperature upon completion of the
temperature sequence as a check on the sample stability. Also, if a
single test specimen has been used over a wide temperature range,
it is desirable to repeat the test on a fresh sample, with a different
initial temperature and following a different sequence of temperatures.

19.2.4.2 Interpretation of Results

Various aspects of the interpretation of rheological data have been


discussed in the previous chapters, and we will try here only to give
some general hints about how to look at results.

584

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

At the least, rheological data tell us something about the processability of the material. Is the viscosity at various shear rates compatible with the requirements of processes such as injection molding or
fiber spinning? Can the process be run in a temperature range
accessible to available equipment? If the answers to these or
related questions are negative, the relations summarized in Chapter
10 can be used to suggest what changes in the polymer structure are
likely to bring the rheological behavior into the required range.
For quantitative analysis a reasonable first step is to examine the
temperature dependence of the viscosity. If the polymer has a low
molecular weight, all of the viscosity data may be near the zero
shear rate value. In this case one can calculate only the activation
energy for viscous flow. If the flow curves extend well into the shear
thinning region it is possible to see if they can be superposed by the
time-temperature superposition principle discussed in Section 2.12.
If superposition is possible, one can obtain a reduced 1]'(w) curve
covering a wider frequency range than is measurable at anyone
temperature.
The curve of 1](-Y) or 1]'( w) should vary smoothly with temperature if the melt is that of a homogeneous and isotropic polymer. If
the Tg of the polymer is known one can test whether the activation
energy EA approximates that calculated from Equation 10-16, which
is the WLF Equation with universal constants. Alternatively, one
can calculate the WLF parameters by the procedure given by Ferry
[15] and compare them with his tabulated values for a variety of
polymers. A departure from the normal range of WLF parameters
would suggest that the sample is either inhomogeneous or
anisotropic.
A very large change of viscosity over a narrow temperature range
may indicate that a phase change has occurred. An example of such
a phase change is the melting of a few imperfect and difficult to
detect crystallites, which act as crosslinks in the melt. Such a change
is ofteh accompanied by a drastic change in the shape of the flow
curve, as shown by polyvinyl chloride for example [16]. Other phase
changes, such as a change from a miscible to an immiscible mixture,
will affect the flow curve similarly [17]. Another possible indication
of such complexities is the failure of the Cox-Merz rule.
Heterophase structure can also cause the melt to exhibit a yield
stress. Concave-upward curvature of a plot of log 1]' versus log w,

INDUSTRIAL USE OF RHEOMETERS

585

with a slope approaching -1 at low frequencies, is an indication of


a yield stress. If a yield stress is suggested by such a plot, it should
be confirmed by measurement of the incomplete recovery of the
stress from a steady shear experiment. A more conclusive test for a
yield stress can be made by use of a constant stress rheometer. A
yield stress also inhibits extrudate swell.
In the absence of such complications, the flow curve can be
analyzed by fitting the data to one of the equations suggested in
Chapters 4 and 10. If elasticity and viscosity data are both available,
it is useful to check their consistency with Graessley's suggested
relation (Equation 10-10) between the compliance and the reduction of the viscosity due to shear thinning. If Mw is known even
approximately, application of Equations 10-5 and 10-7 to the compliance will give an estimate of the MWD of the polymer. Alternatively, Equations 10-11 and Figure 10-7 can be used for such an
estimate. In the absence of any molecular weight information the
power-law slope will give a very rough estimate of Mw/Mn if the
assumptions leading to Figure 10-7 are approximately valid.
Unfortunately, none of these analyses give any direct information
regarding the linearity or branching of the polymer. Comparison
with samples of the same chemical structure that are known to be
linear is required, but this may not be practical. An abnormally high
activation energy and failure of time-temperature superposition
may suggest the occurrence of long chain branching, but only if
other possible reasons for these observations can be ruled out by
independent studies of the structure.
To summarize, we are fairly confident that we understand the
temperature and frequency dependence of the rheology of linear,
isotropic, homogeneous polymer melts. Any departure from expectations is an indication that one of these conditions is not met.
Rheology can take us only so far; beyond that, additional information is required to characterize the system.
19.3 RESIN SELECTION AND OPTIMIZATION AND PROCESS
PROBLEM SOLVING

For a resin user, rheological testing is a valuable tool in the


selection or specification of a resin for a particular application and
for the optimization of processing conditions. For the resin manu-

586

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

facturer the principal application of rheology is in the optimization


of resin formulations for processing, while for a machinery manufacturer it is in the design of equipment for optimal processing of a
class of resins. And all parties are involved in solving problems that
arise when a resin is in use.
The end use requirements generally limit the chemical structures
of polymers to be considered for a given application. Typical
requirements include stiffness, chemical resistance, heat distortion
temperature, impact strength, and dimensional stability. It is then
necessary to identify specific formulations that can be processed
economically, and this is where the rheological behavior is a determining factor.
The problem to be solved must be defined in terms of specific
goals. The goal might be to extrude a smooth-walled tube at so
many feet per minute at some specified maximum pressure, or it
might be to fill a mold in a given cycle time, without exceeding a
specified maximum temperature, or to extrude a parison with a
desired wall thickness profile.
The starting point for these applications is the set of relations
between process behavior and rheological properties that have been
described in the earlier chapters. The use of these relations is
straightforward in principle, but in practice their use may require
anything from simple arithmetic to computation using a sophisticated process simulation.
It may be helpful to give some specific examples. Consider the
limitations on melt viscosity in the melt spinning process. An upper
limit on the viscosity is imposed by the ability of the spinneret
assembly to resist the pressure drop required for flow. It is not
practical to increase the rigidity of the assembly by making it
thicker because of the difficulty and expense of manufacturing very
long holes of small diameter. Neither is it feasible to decrease the
pressure drop by increasing the spinneret diameter, because the
final fiber diameter is fixed by end use requirements. Compensating
for a larger spinneret diameter by increasing the melt drawdown is
not acceptable, because the drawdown affects end use properties
such as fiber modulus and shrinkage. And reducing the pressure
drop by reducing the flow rate is undesirable for economic reasons.
There is also a lower limit. If the viscosity is too low, the speed of
gravity flow, due to the weight of polymer in the spin line, will

INDUSTRIAL USE OF RHEOMETERS

587

exceed the velocity of the extrudate at the spinneret exit, and it will
not be possible to maintain stable operation. Thus, it is not difficult
to set upper and lower bounds for the viscosity for this process,
given the relevant constraints.
The preceding example involved only the use of the fully developed capillary flow equations. However, because the L / D ratio of
spinneret holes is small, it is necessary to consider also the contribution of the entrance pressure drop if the pressure calculated from
the fully developed flow equation approaches the maximum allowable value. Otherwise there are no empirical or arbitrary factors to
take into account.
As an example of a somewhat more complicated problem, consider how to determine whether a simple mold of certain dimensions can be filled by the injection of melt at some pressure and
temperature. An adequate model of the rheological behavior for
this process is usually provided by the power law viscosity equation,
and for this problem one needs to know also the thermal characteristics of the material. One can then answer the question by use of
the relatively simple equations governing freezing off in injection
molds [18]. However, this analysis contains an empirical parameter,
and to use the analysis with confidence it is necessary to determine
its value for a specific polymer by use of spiral flow tests.
The next higher level of complication is encountered when one
considers a process such as extrusion. In Chapter 14 we discussed
the application of numerical simulation methods, which require
rheological data as inputs. These programs can be used to specify
the rheological properties required to achieve an objective defined
by the user. However, it is again important to remember that
absolute prediction by these models is very uncertain; they are most
useful if they have been "calibrated" against known cases.
These examples illustrate the fact either that virtually all commercial processes are too complex to be described adequately by
mathematical equations, or that the equations cannot be solved
with sufficient accuracy for predictive purposes. The strategy practiced, then, is to use the process equipment as a test instrument by
exploring the response of samples of different rheology to the
operating variables of the process.
Another aspect of commercial processing frequently encountered
is that one must meet not just one objective but several. Further-

588

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

more, an alteration of a process variable or of a material property


that favors one objective may move away from another. A strategy
is needed that defines a combination of process and material
parameters that meet all of the requirements, even though the
mechanisms underlying some of the phenomena may not be fully
understood.
Such a strategy is diagrammed in Figure 19-3. The first step is to
determine how a variety of resins that have widely varying rheological properties process over the range of conditions available. A
statistical design of experiments is essential if there are many
independent process variables to adjust. The dependent variables,
or responses, are expressed quantitatively, if possible, or in terms of
a subjective rating scale based on observation. This allows the use
of a statistical analysis to represent the responses as functions of
the process variables.
The second step is to define the allowable limits on each of the
dependent variables. The processability of every resin can then be
characterized by the n-dimensional volume in the space of the n
independent processing variables that is included within these limits. The objective is to find a resin that meets all the requirements.
If none of the available materials is acceptable, it is necessary to
test a wider range of materials. For example, a resin supplier can
modify the polymerization process to optimize the resin formulation
to produce new candidate resins. In principle the procedure described above could be continued until successful. In practice the
use of actual processing machinery for screening tests is time-consuming and expensive, and it may be out of the question if new
variants are produced only in small quantities. The next step, then,
is to analyze the mechanical basis of the dependencies of the
responses on the process variables, in order to relate them to
rheological properties that can be measured in small scale laboratory tests. If a process response is well understood, a formal
analysis may already be available. More likely, a combination of
observation, intuition, and hypothesis based on the shapes of the
response functions is used to define the laboratory tests to be used
for prediction of the processing behavior.
After the appropriate laboratory tests have been proposed, the
obvious next step is to verify their ability to indicate processability.
If they do not, it is necessary to think some more about the

INDUSTRIAL USE OF RHEOMETERS

589

DETERMINE RESPONSE
OF RESINS TO
PROCESSING VARIABLES

ANALYZE
PROCESS
MECHANICS

DEFINE LIMITS OF
ACCEPTABLE RESPONSES

DEVISE
LABORATORY
TESTS
NO

NO

YES

TEST NEW
FORMULATION

DEVISE
QUALITY CONTROLI-+--"
TESTS

SCALE-UP

Figure 19-3. Schematic diagram illustrating strategy for resin optimization.

mechanics of the process. When a series of tests appears to be able


to correlate well with the processing results, they can be used to
screen new variants. The tests may turn out to involve the measurement of well defined rheological properties. In this case the property-structure relations summarized in Chapter 10 can be used to
guide the formulation effort. In other cases some of the laboratory

590

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

tests may be empirical procedures that do not yield values of


specific physical properties. Examples of such tests are extrudate
swell, capillary flow end correction, melt fracture, and melt strength.
In these situations it is important that the conditions of the laboratory test simulate as closely as possible the conditions in the actual
process.
Once formulations that should process acceptably have been
identified, the final step is to evaluate the actual processability. If
the changes in formulation during resin development have been
sufficiently drastic, it may turn out that the predictions based on
laboratory tests are not accurate enough. The new processing data
can then be combined with the original information and the statistical analysis repeated. This may lead to new functional relationships
or to new correlations of the process ability with the laboratory tes.t
results, or even to ideas for improved tests. The entire. formulation,
testing, scale-up, and processing sequence is then iterated until it is
successful, or, less happily, until one is convinced that no solution is
possible.
An application of this approach has been desc.libed in the literature and is summarized here as a concrete example [19]. The
problem involved the parison formation stage of a blow molding
process for making containers with handles by an intermittent, high
shear rate process. The container wall thickness, and therefore the
weight of the container, was specified; if too thin, the container was
not rigid enough, if too thick, it was uneconomical. The parison
diameter had to be large enough to trap a portion of the parison in
the mold in order to form the handle. Two defects of the parison
that had to be controlled were fine scale transverse roughness and
the formation of longitudinal folds or "pleats." The severity of
these defects was evaluated visually and rated on numerical scales.
The only operating variables studied were the discharge (or shot)
pressure and the gap at the exit of a conical bushing-and-mandrel
die. The length of the parison was kept constant.
A large number of commercial and experimental resins were
evaluated at four process conditions, in a two-factor, two-level
factorial design. The results for all the resins were combined into a
single statistical analysis. This made it possible to make sufficiently
accurate predictions even though each resin was tested at only four
conditions. Because there were only two independent variables in

INDUSTRIAL USE OF RHEOMETERS

591

W
II:
:::I

rn
rn
w

II:

a..

b
1020
:I:
rn

DIE GAP (mils)

Figure 19-4. Contour plot for parison diameter. Numbers in squares are measured values of
diameter at corresponding shot pressure and die gap; numbers in circles correspond to lines
calculated from regression analysis of combined data. Adapted from Ref. 19. Copyright
1975 by The Society of Rheology. Reprinted by permission of John Wiley & Sons, Inc.

this particular problem, it was possible to show the functional


dependence of each response in the form of easily interpreted
contour plots. A typical such plot, for parison diameter, is shown in
Figure 19-4.
As the second step of the strategy the allowable limits for each
response were defined. These were superimposed, for each resin,
on an "operating diagram." Figure 19-5 shows the operating diagrams for a commercially acceptable resin and for an unacceptable
experimental resin. The difference in the "window" of operability is
obvious.
The next step was to analyze the dependence of each response on
the process variables to arrive at an understanding of the physics of
the process and to suggest appropriate laboratory tests. For example, from the dependence of parison diameter on die gap and
pressure, and from literature data on parison formation in other
processes, it was hypothesized that extrudate swell was an important parameter. The proposed laboratory test was thus a measurement of the capillary extrudate swell at a shear rate comparable to
that experienced by the melt in parison formation. The swell was
measured at a time after extrusion close to the time scale for the
commercial process.

592

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING


(a)
1200
~

'u;

,9;
LJ.J

a::

::)
(/)
(/)

LJ.J

1100

a::

0..

I-

5 in.

0
I

(/)

1000
5

13

21

17

25

DIE GAP (mils)

(b)
1200~~~~~/-/~----~--~------~

5iV/~
~

..-"'/
PLEATING

::) 1100

1000

MELT
FRACTURE

3;/

2+

XR~?NDIA
4.5 in.

13

17

21

25

DIE GAP (mils)

Figure 19-5. Operating diagrams for (a) acceptable commercial resin and (b) unacceptable
experimental resin. Length of heavy portion of bottle weight line represents acceptable
operating conditions. Note that conditions for acceptability were relaxed in case b. Adapted
from Ref. 19. Copyright 1975 by The Society of Rheology. Reprinted by permission of
John Wiley & Sons, Inc.

The correlation of each observed response with the results of the


laboratory tests was then evaluated. Figure 19-6a gives an example
for pleating, which is the least well understood phenomenon. It
should be noted that the correlation parameter was a combination
of a material property (melt index), a machine setting (die gap), and
a process parameter (parison drop time) that depended on both the
material properties and the machine settings. As shown in Figure

INDUSTRIAL USE OF RHEOMETERS

(a)

6
(!)

i=

a:
(!)

i=

..
,.-

,,'

0..

"

'

........

MELT
SYMBOL INDEX

-.J

~,

,,"."

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

593

0.4
0.6
0.8

16
4
12
8
100 x (MELT INDEX) x
(DROP TlME)/(DIE GAP)

20

(b)

3~---------------':----~
10 mil, 1000 psi

~
w
~

i=

21-

-,........

0..

oa:

I-

1200

..

...........,;:....-

y.~

..-

~.~.

~..
I

10 mil, 1200 psi

1400
1600
APPARENT VISCOSITY (POISE)

Figure 19-6. Correlations of experimental observations with resin and with machine operating conditions: (a) Parison pleating as a function of resin melt index, die gap and parison
drop time; (b) Parison drop time versus high shear rate viscosity. Adapted from Ref. 19.
Copyright 1975 by The Society of Rheology. Reprinted by permission of John Wiley &
Sons, Inc.

19-6b, an auxiliary correlation was used to predict the drop time


as a function of another material property, the high shear rate
viscosity.
The laboratory tests were used to characterize the properties
of additional experimental resins. The understanding of the relation
of the molecular structure and rheology and the rheological bases
of the tests led to successful new formulations that were blends of
polymers from two different process streams [20].

594

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Several comments based on experience with this strategy are in


order.
1. The fortuitous restriction to two independent variables made
possible the representation of the data on two-dimensional
contour plots. When there are more variables, numerical techniques or more complicated plots are required to define the
operating windows.
2. Although the successful application of the strategy is demonstrated in this example, there was still a good deal of empiricism. The relation of diameter and thickness swells was
ignored, for instance. The result was that there were two
polymers, fortunately both acceptable, that behaved similarly
in all of the laboratory tests but had slightly different processing characteristics. Had the difference been commercially significant, a more rigorous analysis or different tests would have
been required.
3. Because of the complex rheological basis of some of the tests
used, it was found advantageous to explore their relation to
molecular structure by determining the properties of mixtures
of components with known properties, as described by
Friedman and Porter [21], despite the non-uniqueness of these
blending rules [22]. Blending rules are also useful in making
formulations that are blends of existing polymers.
Trouble-shooting problems that arise during the operation of a
process has many similarities and some differences, compared to
resin selection and optimization. The similarities arise from the
same need to understand the effects of the deformation and temperature history of the process on the rheological properties of the
polymer. A significant difference is that one is starting from a
known baseline of suitable resin and established processing conditions. Another difference is that the rheologist involved in problem
solving often does not have control of the operation of the process.
He or she may be told that the source of a problem is a change of
resin behavior when it may actually be an inadvertent change in the
process.
On the other hand, there are instances of resin changes that are
not detected by routine quality control (QC) tests. This happens

INDUSTRIAL USE OF RHEOMETERS

595

when the QC tests that are being used are inappropriate. This
question is discussed at some length in Section 19.4. Obviously a
more thorough QC procedure is called for in such a case.
A situation that requires an understanding of the process principles in addition to the melt rheology is one in which a processing
problem occurs although there has not been any change of the resin
molecular structure. Resin tests obviously will not point to the
cause for the process upset. However, the interaction of the resin
properties with the processing conditions may have caused a variation in melt behavior.
To illustrate some of these points, consider the extrusion of a
sheet from a coat-hanger die. Uniformity of the thickness across the
width of the sheet is usually achieved by setting the gap at the die
lips. However, these settings will produce uniform thickness only
for a specific combination of melt temperature and die temperature
profile, because the heat transfer rate varies with the path of the
melt through the die [23]. A change in the melt temperature for any
reason will thus cause an upset in the thickness uniformity. The
change might arise from a detectable variation in a resin characteristic such as pellet size distribution or pellet crystallinity. Or the
cause might be a process problem such as a faulty temperature
controller, an obstruction in a cooling water line, or a variation in
pellet feed temperature.
It is helpful in the case of a lot-to-lot variation to compare
"retain" samples of "good" lots with the material suspected of
causing the problem. Even better, to diagnose whether a problem
arises from the machinery or from the resin, it is informative to
alternate the suspect material with resin known to have performed
well in the past. However, there are many pitfalls for the uninitiated. For example, Ramamurthy [24] found that in a film application, several hours of operation were required to detect differences
in performance among several resins.
19.4 RHEOLOGICAL QUALITY CONTROL TESTS

The product of every polymerization and compounding operation


inevitably exhibits some variability in quality. The purpose of QC
tests is to assure the resin manufacturer and the processor that the
properties of the material produced fall within acceptable produc-

596

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

tion limits. The design of appropriate QC tests and the establishment of production limits is an integral part of product and process
development. The success or failure of a commercial process often
hinges on how well this is done.
The requirements for a QC test are as follows:
1. The test must be sensitive to the property requiring control.
2. The test must be reliable, quick, and economical, and plant
personnel must be capable of carrying it out.

The first of these requirements is absolutely essential. The property


to be controlled may be a measure of processing behavior, or it may
be a rheological property characteristic of molecular structure that
is known to correlate with an end use property. Selection of the
rheological parameters to control therefore requires knowledge of
the structure-processing-property relationships.
The second requirement arises from the fact that QC tests are
run routinely. The operators are often plant personnel, and they do
not generally have the same level of training as laboratory technicians. Finally, the apparatus must be rugged and relatively inexpensive.
The requirements of QC tests are different in some respects from
those of the rheological tests used for resin characterization and
optimization. In general the latter are used to measure well defined
material properties. The result of such a measurement will be
independent of the specific instrument used. A QC test result, on
the other hand, may be affected by several rheological properties.
This is acceptable if only one of these properties is affected by a
change in the polymer production process. Any variation of this
property will then be detected by the QC test. There is, however, a
danger in relying on a test that is affected by a number of properties. The production process may change in such a way that several
properties are affected simultaneously, and the result can be a
compensation of opposing effects, which will upset the correlation
of a QC test result with the property of interest.
The ASTM D1238 Melt Flow Test is perhaps the most commonly
used QC test of a rheological nature, and it nicely exemplifies the
above general discussion. It meets the requirements well. The tester
is relatively inexpensive, the measurement is performed quickly, the

INDUSTRIAL USE OF RHEOMETERS

597

instrument is relatively easy to clean and to maintain, and its


precision is comparable to that of much more complex and costly
rheometers. The ease of cleaning and the precision are in part due
to the rather large capillary diameter and short length. The major
defect of the instrument as a rheometer for the measurement of
viscosity is the result of this capillary geometry. The LjR ratio is
only 7.65, and this is not far from the magnitude of the Bagley end
correction for many polymers. Therefore, the melt flow index depends not only on the viscosity of the polymer but also on the end
correction. If the structure factors controlling both of these parameters can be varied by the polymer production process, and if the
properties to be controlled are functions of these, the melt index
cannot be relied upon to assure the constancy of the properties. It
must then be supplemented by some other test that responds
differently to the structural change. In addition, it is not an apparatus that lends itself readily to automation.
The D1238 test can be employed with different loads in order to
obtain an estimate of the shear rate dependence of viscosity. This is
particularly likely to be useful when it is necessary to control a
property such as impact strength, which is sensitive to the number
average molecular weight Mn' The flow rate at low shear stress is
influenced primarily by the weight average molecular weight, and
the molecular weight distribution influences the shear rate dependence. By checking that both quantities are under control one can
be reasonably confident that the Mn is also controlled.
Additional information that is often useful for QC purposes can
be obtained by measuring the extrudate diameter, either with a
micrometer or by weighing a measured length. Extrudate swell is
probably the most easily measured property that is related to the
compliance, but it is useful only for unfilled polymers. When
making this measurement it must be kept in mind that the swell
depends on die geometry, shear rate, and the time allowed for
swelling to occur before solidification of the extrudate.
Instruments to measure dynamic viscosity, although expensive
compared to a melt index tester, may be justified as QC instruments
for applications in which close control of viscoelastic behavior is
essential. The use of a computer for control, data acquisition and
data reduction makes such an instrument particularly useful for QC
work. If this investment is justified, it will also be worthwhile to

598

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

provide a facility for rapid, reliable sample preparation, because


poor sample preparation is a frequent source of error.
The extensional viscosity is the most difficult rheological property
to measure and control. For QC test purposes the best choice may
be to measure the capillary end correction and to approximate the
extensional viscosity by Cogswell's analysis of converging flow [25]
using equations given in Chapter 8. Somewhat less convenient, but
still usable for QC purposes, is the non-isothermal spinning, or
"melt strength," test. Like the converging flow test, it is affected by
a number of rheological parameters, so that the constancy of the
melt strength is no guarantee that the extensional viscosity is
constant [26]. Nevertheless, it has been found useful as an empirical
characterization parameter for a number of processes. In addition,
it can be used as a direct measure of the phenomenon of draw
resonance, which can be a rate limiting factor in melt spinning and
film casting [27].
Laun and Schuch [28] have measured the true extensional viscosities of polystyrenes and of linear and branched polyethylenes and
have compared these with the estimates obtained from the converging flow and melt strength tests. For the converging flow results,
there was agreement for some resins over narrow ranges of strain
rate, but for other resins there was no quantitative correlation
between the two functions. However, in comparisons of two resins,
the values of the dPent for fixed values of Q and YA were found to
give the same ranking of the materials as the values of the true
extensional viscosity. Laun and Schuch [28] also found that there
was a rough correspondence between the extensional viscosity calculated from the melt strength (see Section 6.3.2) and the true
value.
The ability of these tests to rank materials on the basis of their
extensional flow behavior is of significant interest. The most primitive estimate of the extensional viscosity is the Newtonian fluid
value, which is three times the shear viscosity. However, for polymers that exhibit extension thickening, this estimate can be off by
an order of magnitude. Both the converging flow and the melt
strength test appear to be capable, at least for some polymers, of
distinguishing differences in extensional flow behavior among several resins. For example, Laun and Schuch [28] compared polymers

INDUSTRIAL USE OF RHEOMETERS

599

whose shear viscosIties were nearly indistinguishable but whose


extensional viscosities were measurably different. Although one
must be cautious in the interpretation of the data, it appears that
there are situations in which the entrance pressure drop or melt
strength can be useful for QC testing, especially when variations in
elongational viscosity are important.
To recapitulate, QC tests are best chosen on the basis of an
understanding of the mechanics of the process, of the dependence
of rheology on resin structure, and of the variability of structure
with the polymerization and formulation process. When empirical
tests are used, they should be run under conditions approximating
those in the process, if possible. Properties other than rheological
ones that can affect the process must not be overlooked.
Finally, it must be remembered that the test results depend
crucially on maintaining the instrument in good order, in keeping it
calibrated, and on ensuring that operators run the test correctly.
Absolute calibration should be done periodically, with a frequency
indicated by experience in the particular environment. Standardization, by measurement of one or more lots of resin kept specifically
for this purpose, should be done frequently, perhaps even daily.
Any significant variation is a warning to be followed up by a search
for its cause. It is useful to keep a log book of the standardization
test results, because a drift of results with time can indicate the
need for calibration or adjustment more clearly than a single
measurement can.
REFERENCES
1. M. Mooney, "Symposium on Consistency," pp. 9-12, American Society for

2.
3.
4.
5.
6.

Testing Materials, Philadelphia, June 29, 1937. For a more extensive quotation, see Appendix IV, "Experimental Methods in the Scientific Mechanical
Testing of High Polymers" in T. Alfrey, Jr., Mechanical Behavior of High
Polymers, Interscience, New York, 1948.
J. M. Dealy, Rheometers for Molten Plastics, Van Nostrand Reinhold, New
York, 1982.
A. Rudin and H. P. Schreiber, Polym. Eng. Sci. 23:422 (1983).
K. F. Wissbrun and A. C. Griffin, 1. Polym. Sci. Polym. Phys. 20:1835 (1982).
R. M. McGlamery and A. A. Harban, Mater. Res. Stand. 3, 12:1003 (1963).
S. Wu, Polym. Mater. Sci. 50:43 (1984); Polym. Eng. Sci. 25:122 (1985).

600

7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

W. H. Tuminello, Polym. Eng. Sci. 26:1339 (1986).


M. Kimura, G. SaIee and R. S. Porter, J. Appl. Polym. Sci. 29:1629 (1984).
R. W. Lenz and A. N. Schuler, J. Polym. Sci. Polym. Symp. 63:343 (1978).
E. L. Lawton, Polym. Eng. Sci. 25:348 (1985).
J. Horrion, R. Jerome, Ph. Teyssie, C. Marco and C. E. Williams, Polymer
29:1203 (1988).
K. F. Wissbrun and A. C. Zahorchak, J. Polym. Sci. A-J 9:2093 (1971).
S. A. Jabarin and E. A. Lofgren, Polym. Eng. Sci. 26:620 (1986).
J. L. White and H. Yamane, Pure Appl. Chem. 57:1441 (1985).
J. D. Ferry, Viscoelastic Properties of Polymers, Third Ed., John Wiley & Sons,
New York, 1980, p. 273.
L. A. Utracki, Z. Bakerdjian and M. R. Kamal, Trans. Soc. Rheol. 19:173
(1975).
A. Ghijsels and J. Raadsen, Pure Appl. Chem. 52:1359 (1980).
A. M. Hull, S. M. Richardson and J. H. Selopranoto, Plast. Rubber Proc.
Applic. 6:183 (1986).
J. S. Schaul, M. J. Hannon and K. F. Wissbrun, Trans. Soc. Rheol. 19:351
(1975).
J. S. Schaul, K. F. Wissbrun and M. J. Hannon, U.S. Patent 3,784,661, Jan. 8,
1974.
E. M. Friedman and R. S. Porter, Trans. Soc. Rheol. 19:493 (1975).
K. F. Wissbrun, Trans. Soc. Rheol. 21:149 (1977).
B. Vergnes, P. Saillard and J. F. Agassant, Polym. Eng. Sci. 24:900 (1984).
A. V. Ramamurthy, J. Rheol. 30:337 (1986); Adv. Polym. Technol. 5:489
(1986).
F. N. Cogswell, J. Non-Newt. Fluid Mech. 4:23 (1978).
K. F. Wissbrun, Polym. Eng. Sci. 13:342 (1973).
W. K. Lee, Chem. Eng. Commun. 53:117 (1987).
H. M. Laun and H. Schuch, J. Rheol. 33:119 (1989).

Appendix A
Measures of Strain for Large
Deformations
A.1 THE DISPLACEMENT FUNCTIONS AND THE DISPLACEMENT GRADIENT
TENSOR

As the first step in establishing a quantitative measure of the strain


that occurs in a fluid as a result of a large deformation, we need to
describe the position of a particle of fluid as it moves about during
a deformation process. This is easily done by writing down the
coordinates of the particle as functions of time. However, we need
to keep track of which particle we are tracking. This can be done by
"labelling" each particle with its position vector, x, at some reference time, t 1 Thus, giving the components of x(t 1 ) identifies a
particular fluid particle and distinguishes it from all other particles.
At some other time, t 2 , the coordinates of this particle will be given
by the following "displacement functions":
t 2' x{t 1)]
X2 =X 2[t 2 ,X(t 1)]

(A-Ia)
(A-Ib)

X3[t 2,X{t1)]

(A-Ic)

Xl = Xl [

X3

These functions tell us the location at any time t 2 , of the particle


that is located at X(t1) at time t 1
As an example of the use of the displacement functions to
describe a deformation, consider simple shear flow in which the
shear strain, 1', is some specified function of time, y(t), with
1'(0) = O. The displacement of a typical fluid particle is shown in
Figure A-I. We locate the origin for the position vectors of all fluid
particles at point 0 on the lower, stationary plate. Since we shall be
considering only one or two specific particles in the remainder of
601

602

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

X,(t2)-j
X,

(t'):::;:::i

X2(t,)

o
Figure A-I. Fluid particle displacement in simple shear.

our discussion, it will not be necessary to include the particle label,


X(t1)' as an argument of the displacement functions. Thus, the shear
strain can be expressed in terms of the displacement functions as
follows for the particular particle shown:
y(t2) - y(tl)

x j (t 2) - XI(t l )
X2(t l )

(A-2)

where [X I(t 2 ) - xj(t l )] is the distance moved in the XI direction by a


fluid particle that is at a distance x 2 from the stationary plate,
during the time interval between t1 and t 2 . The displacement
functions for this deformation can then be expressed in terms of the
shear strain as follows:
Xj

(t 2)

X 2(t 2 )
X3

x 1(t j ) + X 2(t 1)[ y(t2) - y(t1)]

= X 2(t l )

(t 2 ) =

X3

(A-3a)
(A-3b)

(t l )

(A-3c)

To describe the deformation of a fluid element, we need to


examine the relative displacement of two fluid particles. Consider
two neighboring particles that are separated by the vector dx(t I) at
time tl and by dX(t2) at time t 2, as shown in Figure A-2. If we
could define a quantity relating these two vectors, it might be useful
to describe the strain that has occurred between times tl and t 2
One such a quantity is the "displacement gradient tensor,"
F;/tl' t 2), which is defined as follows:
Fij (t p t 2 ) ==

3x;{t 2 )

()

3xj tl

(A-4)

APPENDIX A

603

t--I,-----IlX,-------l
,-------,

t,

o
Figure A-2. Position and displacement vectors for two particles of an element of a body of
fluid undergoing solid body translation.

Another tensor that relates the two vectors dx(t 2 ) and dx(t l ) is the
"inverse" of the displacement gradient tensor, which is defined as
follows l :
(A-5)
The use of the displacement gradient tensor to determine dx(t 2 ),
given dx(t l ), is demonstrated below:
dx j (t 2 )

Fil(t l , t 2 ) dxl(t l )

+ Fi2 (t!> t 2 ) dx 2 (t l ) + Fi3 (t l , t 2 ) dx 3 (t l )


(A-6)

The inverse tensor can be used in a similar manner to determine


dx(t l ) given dx(t 2 ).
lNote that the inverse of a tensor is not the tensor whose components are the reciprocals of
the components of the original tensor.

604

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

To see if these tensors are useful as measures of strain, we


consider first the "solid body translation" of the fluid element
shown in Figure A-2. In this type of motion, there is no deformation of the fluid element, i.e., no strain. To see if the displacement
gradient tensor or its inverse reflect this, we first write the displacement functions for this motion:

Xl(t l ) + dX 1

(A-7a)

X2(t2) = X2(t1) - dX2

(A-7b)

XI(t2)

X3

(t 2 ) =

X3

(t 2 )

(A-7c)

Now we want to evaluate Fij and its inverse for this flow. We note
that the two vectors dx(t 2 ) and dX(tI) are equal; this implies (from
Equations A-4, A-5, and A-7) that the components of F;/t l , t 2 ) and
its inverse have the special, simple form shown below:

o
1

~l

(A-8)

where the quantity on the right is called the "unit tensor." When a
vector is multiplied by the unit tensor it is unchanged, and since

t = t,

o~-------------------x'

o~------------------x,

Figure A-3. Displacement vectors for two particles of an element of a body of fluid
undergoing solid body rotation.

APPENDIX A

605

there is no strain in this example, that is exactly the way we want a


strain tensor to behave. Thus, both FJt l , t 2) and its inverse show
promise as possible measures of finite strain.
As a further test of the usefulness of these tensors as measures of
strain, consider the "solid body rotation" shown in Figure A-3.
Again there is no deformation, i.e., no strain. However, in this case
the vectors dX(t2) and dx(t l ) are not equal and neither Fi/t l , t 2)
nor its inverse reduces to the unit tensor. Thus, neither of these
tensors is, after all, a useful measure of finite strain.
A.2 THE CAUCHY TENSOR AND THE FINGER TENSOR

There are other tensors, however, that do have the desired properties and are thus useful as measures of strain for large deformations. Two of these are the Cauchy and Finger tensors, which are
defined as follows:
(A-9)
(A-10)
A.3 THE DISPLACEMENT GRADIENT TENSOR FOR SIMPLE SHEAR AND SIMPLE
EXTENSION

In order to derive the components of the Cauchy and Finger


tensors for simple shear and simple extension (Equations 3-1 to 3-4)
it is useful to have available the components of Fij and its inverse
for these two deformations. For simple shear flow the components
of these tensors are:

~l
~l

(A-H)

(A-12)

The components of the Cauchy tensor for simple shear (Equation


3-1) are obtained by combining Equation A-9 with Equation A-H,

606

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Figure A-4. Principal component of the position vector for simple extension of a sample with
one end fixed at Xl = O.

while the components of the Finger tensor (Equation 3-2) are


obtained by combining Equation A-lO with Equation A-12.
For simple extension, with x I taken to be the principal stretch
direction (see Figure A-4), the distances of a fluid element from the
Xl = 0 plane at times tl and t2 are related to the Hencky strain, e,
evaluated at these two times, by:
(A-13)
The displacement functions are:

X1(t 2) = xl(tl)exp[ e(t 2) - e(t l )]

(A-14a)

X2(t 2) = x 2(t 1)exp{( -1/2)[ e(t2) - e(t l )]}

(A-14b)

X3(t 2) = x 3(t l )exp{( -1/2)[ e(t 2) - e(t l )]}

(A-14c)

And the components of the displacement gradient tensor and its


inverse are:

o
e{( -1/2)[E(t2)-E(tj)]}

o
(A-IS)

(A-16)

Appendix B
Molecular Weight Distribution
and Determination of
Molecular Weight Averages
Synthetic polymers are not composed of chains of equal lengths.
Rather, they have a distribution of lengths or molecular weights.
The exact form of the distribution depends upon the details of the
polymerization mechanisms, and may also be affected by subsequent treatment. Treatment with solvents can selectively remove
low molecular weight fractions. Shear induced degradation can
preferentially cleave the higher molecular weight components.
Chemical reactions such as oxidation can cause either chain scission
or crosslinking. Fractionation according to molecular weight can
occur during complete precipitation from a solution, so that the
average molecular weight of a polymer in powder form may depend
on the particle size distribution of the powder.
Various physical properties, including rheological properties, have
different dependencies on molecular weight distribution (MWD),
and it is therefore important to be able to describe and measure the
MWD. Conversely, the determination of the MWD curve is done by
measuring properties that have a known dependence on molecular
weight. The most commonly used techniques include light scattering from dilute solution, dilute solution viscosity, dilute solution
osmotic pressure, and rate of elution from a porous medium (gel
permeation or size exclusion chromatography-GPC or SEC) [1,2].
A distribution can be described by specifying the weight fraction
of the polymer that has a given molecular weight. (It can also be
specified by the number fraction of a given molecular weight; the
two specifications are easily interconverted). Such a specification
607

608

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

5.0

4.0

'"o,....
X

B3.0
~
II:
L1.

a
2.0
iii

I-

1.0

o.o~~

10

__. .

..

~~~~~~~

100

1000

~~~~~~

10000

__

~~~

1 ooooq

1 000000

MOLECULAR WEIGHT

Figure B-l. Differential molecular weight distribution curves for most probable (solid curve)
and log normal (dashed curve) molecular weight distributions, both with Mw/Mn = 2.

can be presented in the form of a diagram like that shown in Figure


B-l. It is usually more convenient to represent the MWD by a
smooth curve that can be represented by a mathematical function.
Often such a function can be derived from a knowledge of the
kinetics of the polymerization. It is most convenient if this function
contains only 2 or 3 parameters, which can then be expressed in
terms of the commonly measured average molecular weights.
For any distribution function one can define various averages.
For example, if the polymer contains a weight fraction wi(M) of
polymer of molecular weight M j , the weight average molecular
weight Mw is defined as

=
w

IMw.(M.)
Iwj(MJ
I

(B-1)

APPENDIX B

609

By definition of the weight fraction, the denominator of Equation


B-1 is unity. The summation extends over all possible values of the
index i. It should be noted that the distribution wi(M) is often
expressed in the literature as a continuous function w(M) dM, the
weight fraction of molecules of molecular weight between M and
M + dM. In that case, and similarly for the other averages discussed below, the summation is replaced by an integral. To illustrate, for Mw this takes the form

i Mw(M)dM
io w(M)dM
00

Mw

= --"-0-00- - - - -

(B-1a)

The number average molecular weight Mn is defined similarly by


a summation of Mi over the number fraction ni(MJ Since ni(M)
is related to wi(M) by the expression ni(M) = wi(Mi)/Mi, this
average can be expressed in terms of the weight fraction distribution as

(B-2)
Various other higher molecular weight averages can be defined
similarly. Two that are important in rheology are:

(B-3)

(B-4)
In order to calculate the average molecular weight of a blend of
polymers, for each of which the averages are known, it is necessary
to sum the weighted numerators and denominators of these expressions individually [3]. This leads to the following blending rules; the
symbol <Mx) denotes the x-average molecular weight of the blend,
WI and w 2 are the weight fractions of the components of the blend,
and M; and M;' are the x-average molecular weights of compo-

61 0

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

nents 1 and 2, respectively:

As shown diagramatically in Figure B-1, the different averages are


weighted more or less heavily by the low and high ends of the
MWD, the Mn being smallest and Mz+ 1 highest.
As mentioned above, the mathematical form of the MWD depends on the chemistry of the polymerization and any subsequent
treatment. The reader is referred to standard polymer chemistry
textbooks, including the classic one of Flory [4] for details. Two of
the most important MWDs are given here. The "most probable"
MWD has the form:
(B-5)

The successive average molecular weights for this MWD are in the
ratios

The most probable MWD is the one that results from polycondensat ion reactions, such as those for Nylon-6,6 and polyethylene
terephthalate, and for other polymerizations involving a randomization step, as for acetal copolymer.
A more complicated MWD is the "log-normal" distribution,
so-called because the plot of w(M) against the logarithm of the M

APPENDIX B

611

follows the Gaussian normal error curve. It has the form:


(B-6)

Mo characterizes the location of the distribution and /3 its breadth.


The weight average molecular weight M w is

and the ratios of molecular weights are

This MWD is encountered with polymers such as HDPE and


polypropylene that are produced by polymerization with a heterogeneous catalyst. Depending of course on the value of /3, the MWD
can be extremely broad. A typical HDPE can have, for instance, an
Mn of 10 4 , an Mw of 10 5 , an M z of 10 6 , and at least theoretically,
an Mz+ 1 of 10 7
As for methods of determining the MWD, the reader is again
referred to polymer textbooks. However, briefly, Mn is measured by
methods that count the number of molecules, such as osmotic
pressure of dilute solutions or chemical or spectroscopic end group
determinations. Mw is measured by light scattering, which can also
give an indication of M z Undoubtedly the method of choice for
determining the complete MWD is GPC, which does, however,
require calibration, usually with narrow MWD polymers whose
molecular weights have been determined by one of the above
absolute techniques.
A very convenient measure of molecular weight is the dilute
solution viscosity. The quantities usually used to describe the solution behavior are the intrinsic and inherent viscosities, which are
defined in Appendix C. Again, this requires calibration, and if the
polymer of interest has a very different MWD than the calibration
material, the molecular weight average determined will be somewhat in error. Generally, however, it will be close to Mw.

612

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

One other caution is that long chain branching affects both the
GPC elution volume and the intrinsic viscosity. Comparison of a
variety of measurements is therefore required to establish the
existence of and to estimate the amount of long chain branching.
This is by no means a trivial exercise, and should probably be
considered to be more of a research project than a routine
measurement.
REFERENCES
1. A. R. Cooper, Polym. Eng. Sci., 29:2 (1989).
2. S. Balke, Quantitative Column Liquid Chromatography, Elsevier, New York,
1984.
3. K. F. Wissbrun, Trans. Soc. Rheol. 21:149 (1977).
4. P. J. Flory, Principles of Polymer Chemistry, Cornell University Press, Ithaca,
1953.

Appendix C
The Intrinsic Viscosity and
the Inherent Viscosity
C.1 DEFINITIONS OF SOLUTION VISCOSITY QUANTITIES

The most common method for determining an approximate average


molecular weight for a polymer is to measure the viscosity of a
dilute solution [1]. The measurement is usually made using a glass
capillary viscometer.!
Several quantities are used to describe the viscosity of a solution,
and we define these below. When two names are given, the first is
that proposed by Cragg [2], and the second is that proposed by
IUPAC [3].
The relative viscosity or viscosity ratio:

(C-l)
where 'TIs is the solvent viscosity.
The specific viscosity:
'TIsp

==

'TIr - 1

(C-2)

The reduced viscosity or the viscosity number:

(C-3)
where C is the concentration in g/dl (g/lOO mI).
lA vibrational viscometer such as that made by Nametre can, in principle also be used. It is
faster and more convenient but requires a larger sample and is somewhat less precise. It is
also more difficult to control the temperature of the sample.
613

614

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

CONCENTRATION, gjdl

Figure C-l. Sketch showing the general behavior of two functions of the solution viscosity as
functions of concentration. Both converge to the intrinsic viscosity in the limit of zero
concentration.

The inherent viscosity or the logarithmic viscosity number:

(C-4)
This quantity is not to be confused with the "inherent melt viscosity" defined in ASTM D3835 [6]. The latter is the result of an
extrapolation technique designed to eliminate the effects of thermal
degradation in the use of a capillary rheometer to determine melt
viscosity.
The intrinsic viscosity or the limiting viscosity number:

(C-5)
While both the reduced and inherent viscosities approach the
intrinsic viscosity in the limit of vanishing concentration, they diverge at larger concentrations, as indicated in Figure C-l.
For values of [ '17] of about 2 or higher, the solution may be shear
thinning, and in this case an extrapolation to zero shear rate is
required to obtain the intrinsic viscosity [4]. For ultrahigh molecular
weight polyethylene, a special rotational rheometer has been recommended to obtain data for such an extrapolation [5].

APPENDIX C

615

C.2 RELATIONSHIP TO MOLECULAR WEIGHT

The use of solution viscosity to provide information about molecular weight arises from the observation that for linear, mono disperse
polymers, the intrinsic viscosity is related to the molecular weight as
follows:
(C-6)
where K' and a depend on the polymer, the solvent and the
temperature. Extensive tabulations of the constants have been
published [7-9]. Equation C-6 is called the Mark-Houwink-Sakurada
equation.
In the case of polydisperse materials, one can use this relationship to define a "viscosity average molecular weight":
(C-7)

It can be shown that this average is related to the molecular weight


distribution as follows:

(C-8)
When a = 1, Mv becomes equal to Mw' Often Mv is about 80 to
90% of M w , and an often-used approximation of (C-7) is:

[71] =

k'M~

(C-9)

In fact, k' and a in this equation are not truly independent of


molecular weight, and Manaresi et al. [10] have proposed a correction term for (C-9) that depends on M n, Mw and M z.
C.3 RAPID TESTING FOR QUALITY CONTROL PURPOSES

We note that the determination of the intrinsic viscosity requires


measurements of the viscosities of several solutions and an extrapolation to zero concentration. For routine quality control, a quicker
and more common practice is to simply measure the inherent
viscosity (IV.) at some standard value of the concentration, usually

616

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

between 0.1 and 1 gldl and often around 0.5. While this does not
lead directly to a quantitative determination of the average molecular weight, it does provide quick evidence of a variation in Mw from
one sample to another.
REFERENCES
1. F. W. Billmeyer, Jr., Textbook of Polymer Science, Second Edition, John Wiley
& Sons, N.Y. 1971.
2. L. H. Cragg, J. Colloid Sci. 1:261 (1946).
3. International Union of Pure and Applied Chemistry, J. Polym. Sci. 8:255
(1952).
4. B. H. Zimm and D. M. Crothers, Proc. Nat. Acad. Sci. 48:905 (1962).
5. H. L. Wagner and J. G. Dillon, J. Appl. Polym. Sci. 36:567 (1988).
6. American Society for Testing and Materials, "Standard Test Method for
Rheological Properties of Thermoplastics with a Capillary Rheometer,"
D3835-79 (Reapproved 1983), Appendix.
7. M. Kurata and W. H. Stockmayer, Adv. Polym. Sci. 3:196 (1963).
8. M. Kurata, Y. Tsunashima, M. Iwama and K. Kamada, Chapter IV, p. 1,
Polymer Handbook, Ed. by J. Brandrup and E. H. Immergut, Second Edition,
John Wiley & Sons, N.Y. 1975.
9. N. C. Billingham, Molar Mass Measurements in Polymer Science, John Wiley &
Sons, N.Y. 1977.
10. P. Manaresi, A. Munari, F. Pilati and E. Marianucci, Eur. Polym. Joum.
24:575 (1988).

Appendix D
The Glass Transition
Temperature
The temperature and pressure dependencies of the viscosity are
closely related to a characteristic temperature, Tg For the sake of
simplicity, we consider a polymer such as ordinary (atactic)
polystyrene whose structure is not sufficiently regular to permit
packing into a crystalline form. At a very low temperature it is a
rigid solid. If we heat this solid and measure its specific volume (the
reciprocal of the density), we find that it increases, as does that of
all solids, because of the increased thermal motion of its atoms. The
thermal expansion coefficient a is defined as
a

~(

dv )
v dT

(D-l)

where v is the specific volume and T the absolute temperature.


Typically a is on the order of 2 X 10 - 4 K - 1.
As we continue to heat the material we find that in a very narrow
temperature range the thermal expansion coefficient begins to
increase rapidly and becomes two- to three-times greater. This is
illustrated in Figure D-l. The intersection of the straight lines that
approximate the specific volume above and below this temperature
range is called the "glass transition temperature" (Tg). Below Tg
the material is a hard, glassy solid; above Tg it softens visibly and
becomes a viscous liquid.
The Tg is not a sharply defined temperature. Its value depends
on the rate of heating (or cooling) and on the previous thermal
history of the sample. However, these variations are relatively small
(on the order of 3C for a ten-fold change of rate) and we can
consider Tg to be a characteristic of the material. Other thermodynamic quantities, including the specific heat Cp , show a similar
617

618

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

UJ

::i!
:::>
...J

o
>

()

u:::

U
UJ
Q.

(f)

TEMPERATURE

Figure D-l. Glass transition temperature Tg , as inferred from the intersection of extrapolations of the two straight-line segments of the specific volume versus temperature curve.

transItIon. Probably the most convenient way to measure Tg is by


determining the change of Cp by differential scanning calorimetry
(DSC) or differential thermal analysis (DTA).
A number of theories have been advanced to account for the
phenomenon of a glass transition. These are not described here; the
interested reader may consult the literature [1]. Briefly, however,
we can describe what is happening as follows. The volume of the
atoms comprising the material is essentially independent of temperature. The observed increase of the specific volume as the polymer
is heated is ascribed to the introduction of "free volume," vf' When
the free volume reaches a critical value (on the order of 2.5% of the
total volume), the chain segments of the polymer have sufficient
room to move freely on the time scale of the heating experiment.
As a result they are able to respond to a further temperature
increase by increasing vf more rapidly. The ability of the segments
to move freely is reflected in the softening that is observed above

APPENDIX D

619

Tg This motion is observable not only by the mechanical response,

but also by other measurements of mobility, such as dielectric loss


and nuclear magnetic resonance (NMR).
Polymers that have a regular chain structure, such as linear
polyethylene, polypropylene, and polyethylene terephtbalate also
show the glass transition phenomenon. The observation is complicated, however, by their crystallinity, which preserves the solid-like
character of the polymer above Tg Polymers such as PET crystallize
slowly and can be quenched into the amorphous glassy state by
rapid cooling; conventional volume or specific heat measurements
are then possible. Also, if the maximum degree of crystallinity is not
too high, the Tg of the "amorphous fraction" of the polymer can be
observed conventionally. For highly crystalline polymers such as PE
or polyacetal, the Tg is more readily determined by methods sensitive to mobility, such as dynamic mechanical analysis. Since many
polymers show transitions other than the Tg , these methods can
lead to some ambiguity.
The value of Tg depends strongly on molecular structure. It is
increased by substituent groups that stiffen the chain, as is to be
expected from the effect of chain mobility. Polarity, which increases
interchain cohesion, also raises the Tg A few representative examples to illustrate these points are shown in Table 10-3. Extensive
tables are available in the Polymer Handbook [2]. Van Krevelen and
Hoftyzer [3, p. 99] also tabulate data as well as "group additivity
contributions" from which the Tg of an untabulated polymer can be
estimated.
The Tg values listed in Table D-l are for high molecular weight
polymers. The extra free volume associated with chain ends decreases the Tg of low molecular weight polymers according to

(D-2)
For polystyrene K is about 10 5 [4, 5], so that the Tg of a
polystyrene with an Mn of 10 4 is 10C lower than that of a very high
molecular weight polymer.
Plasticizers also have a strong effect on Tg for the same reason,
i.e. they contribute free volume. It should be noted that small
amounts of solvent may be retained tenaciously by polymers and
can cause appreciable lowering of the observed Tg

620

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Table D-1. Variation of Tg with Molecular Structure


STRUCTURE

TiOC)

Polyethylene
Polypropylene

-CHz-CH z-CH -CH-

-125

Polystyrene

CH 3
-CHz-CH-

POLYMER

Polymethyl Acrylate

cf>

-CHz-CH-

-13

+ 100
+10

COOCH 3
CH 3

Polymethyl Methacrylate

I
I

-CH - C -

+105

COOCH 3

The Tg of a random copolymer or of a miscible blend of two


polymers can be estimated by various mixture rules, of which the
simplest is merely the weight or volume averaging of the component
Tg values. Another expression averages the reciprocals,

(D-3)
On the other hand, block or graft copolymers or immiscible
blends will exhibit the Tg's of both components, perhaps modified
by finite block size or by partial miscibility.
Considering the relationship of free volume vf to Tg , it is not
surprising that Tg depends on pressure. The magnitude of this
effect is to increase Tg about 1.5C per 1000 psi (0.025C per 10 6
dynes/cm 2 ). This is about the value one would calculate from the
compressibility of a polymer melt.
Finally, to close this discussion of transition temperatures, we
should point out that although the Tg is the lowest temperature at
which we can consider a material to be liquid, in many cases
polymers solidify at a higher temperature, the melting temperature
Tm' Polymers vary widely in their ability to crystallize, in the
maximum degree of crystallinity attainable from the melt, and in
the sensitivity of their rate of crystallization to temperature and
cooling rate. A rule of thumb applicable to undiluted homopolymers is that Tg is between 0.5 and 0.67 times Tm , where both
temperatures are expressed in absolute (Kelvin) units. However,

APPENDIX D

621

small variations of structure that affect chain regularity, and thus


the ability to pack, can change Tm by large amounts. It is worth
noting that Tm is very sensitive to small quantities of comonomer
but relatively insensitive to diluents, whereas the converse is true
for Tg
REFERENCES
1. A. Eisenberg, "The Glassy State and the Glass Transition," Chapter 2 in

2.
3.
4.
5.

Physical Properties of Polymers, 1. E. Mark, ed., American Chemical Society,


Washington, D.C., 1984.
W. A. Lee and R. A. Rutherford in Polymer Handbook, Second Edition,
1. Brandrup and E. H. Immergut, Eds., Wiley & Sons, New York, 1975.
D. W. Van Krevelen and P. 1. Hoftyzer, Properties of Polymers, Second Edition,
Elsevier, New York, 1976, p. 99.
T. G. Fox and P. 1. Flory, J. Appl. Phys. 21:581 (1950).
T. G. Fox and P. 1. Flory, J. Polym. Sci. 14:315 (1954).

Appendix E
Manufacturers of
Melt Rheometers and
Related Equipment
This list was compiled for the use of readers who wish to obtain
further information about commercial melt rheometers. It was the
intention of the authors to include all companies offering such
equipment, and any omission is the result of oversight and not of
any selection process. Inclusion in this list does not imply any
endorsement or recommendation by the authors. The information
given was verified in 1989 but will naturally become less accurate
with the passage of time.
Bohlin Reologi Inc.
2540 Route 130, Suite 105
Cranbury, NJ 08512
(609) 655-4447
VOR Rotational Rheometer (cone-plate with torque and
normal force measurement)
C. W. Brabender Instruments, Inc.
P. O. Box 2127
South Hackensack, NJ 07606
(201) 343-8425
Plasti-Corder (torque rheometer)
Extrusiograph (laboratory extruder)
Carri-Med Ltd.
Glebelands Centre, Vincent Lane
Dorking, Surrey RH4 3YX
England
622

APPENDIX E

623

North American Representative:


Mitech Corporation
1780 Enterprise Parkway
Twinsberg, OH 44087
(216) 425-1634
CSL Rheometer (constant stress rheometer for creep and
recovery)
Weissenberg Rheogoniometer (cone-plate with torque and
normal force measurement)
Carter Baker Enterprises ltd.
P. O. Box 2
Stansted, Essex CM24 8JG
England
(0279) 814810
ACER Series 2000 Capillary Extrusion Rheometer
Ceast U.S.A. Inc.
P. O. Box 3072
Fort Mill, SC 29715
(803) 548-6093
Rheovis 2100 (capillary rheometer)
Modular Flow Index (melt indexer)
Costech Associates, Inc.
1184 Corner Ketch Road
Newark, DE 19711
(302) 239-2207
Costech 2000 Process Simulator (biaxial inflation rheometer)
Custom Scientific Instruments, Inc.
13 Wing Drive
Cedar Knolls, NJ 07927
(201) 538-8500
Melt Index Apparatus CS127
Melt Elasticity Indexer Model CS245 (concentric cylinder
recoil apparatus)

624

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Daventest Limited

Tewin Road
Welwyn Garden City
Herts AL7 lAQ
England
Davenport Melt Viscosity System (gas-driven capillary
viscometer)
Davenport Melt Flow Indexers
Dupont Company

Instrument Systems
Concord Plaza, Quillen Building
Wilmington, DE 19898
(302) 772-5500
983 Dynamic Mechanical Analysis System
943 Thermomechanical Analyzer
2970 Dielectric Analyzer
Gottfert Werkstoff-PrOfmaschinen GmbH

Postfach 1220
6967 Buchen
Federal Republic of Germany
North American representative:
Goettfert Incorporated

488 Lakeshore Parkway


Rock Hill, SC 29730
(803) 324-3883
Viscotester 1500 (bench-top capillary rheometer)
Rheograph 2002 (high-pressure capillary rheometer)
Melt Indexer
Bypass-Rheograph (on-line capillary rheometer)
Side-Stream Rheograph (on-line capillary rheometer)
Rheostrain (extensional rheometer)
Karl Frank GmbH

Postfach 1320
D-6940 Weinheim
Federal Republic of Germany

APPENDIX E

625

North American representative:


Carl G. Brimmerkamp & Company, Inc.
102 Hamilton Avenue
Stamford, CT 06902
(203) 325-4101
Melt Indexer
High-Pressure Capillary Rheometer
HBI/Haake

Fisons Instruments
244 Saddle River Road
Saddle River, NJ 07662-6001
(800) 631-1369 (In NJ: 201-843-2320)
Rheodrive System
System 90 (torque rheometers and extruder-fed capillary
rheometers)
Instron Corporation
100 Royall Street
Canton, MA 62021

(617) 828-2500
4200 Rheology Test System (bench-top capillary rheometer)
3210 Capillary Rheometer Accessory (for use with universal
testing machine)
3211 Rheological Testing Instrument (capillary rheometer)
3250 Rotational Rheometer (cone-plate with torque and
normal force measurement)
Kayeness, Inc.

Box 30
Honeybrook, PA 19344
(215) 273-3711
Melt Indexer
Galaxy V Capillary Rheometer
Killion Extruders, Inc.

56 Depot Street
Verona, NJ 07044
(201) 239-0200
Laboratory Rheology Extruder

626

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Metravib ftD.S.
BP 182

69132 Ecully Cedex


France
Viscostrain extensional rheometer
Viscoanalyser (compressional oscillation rheometer)
Metrilec Sari
20, Rue Michal
75013 Paris
France
Rheoplast RClO (pre-shear capillary rheometer)

Monsanto Instruments & Equipment


2689 Wingate Ave.
Akron, OH 44314
(216) 745-1641
MDR2000 Moving Die Rheometer (oscillatory shear cure
meter)
ODR 2000 Oscillating Disc Rheometer (for curing studies)
Capillary rheometer
Die swell detector
MTS Systems Corporation
Box 24012
Minneapolis, MN 55424
(612) 937-4000
Computer Controlled Capillary Rheometer
Model 831 Elastomer Test System
Nametre Company
101 Forrest Street
Edison, NJ 08840
(201) 494-2422
Vibrational Viscometer (for low viscosity liquids)

APPENDIX E

627

Polymer Laboratories, Inc.


P. O. Box 1581
Stow, OH 44224
(216) 688-7339

Dielectric Thermal Analyser


Dynamic Mechanical Thermal Analyser
Rheometrics, Inc.

One Possumtown Road


Piscataway, NJ 08854
(201) 560-8550

Mechanical Spectrometer (cone-plate with torque and normal


force measurement)
Dynamic Analyzer (small amplitude oscillatory shear)
Stress Rheometer (rotational constant torque rheometer)
Extensional Rheometer (based on Miinstedt design)
Melt Analyzer (for automatic quality control)
On-Line Rheometer (concentric cylinder rheometer)
Melt Flow Monitor (on-line slit rheometer)
Optical Analyzer (optical accessory)
Rosand Precision Limited
11 Little Ridge

Welwyn Garden City


Herts AL720H,
England
Rosand Automatic Melt Indexer
Seiscor Technologies
P. O. Box 470580
Tulsa, OK 74147-0580
(918) 252-1578

CMR-II Process Rheometer (on-line capillary rheometer)


Flow Characterization Rheometer (on-line capillary rheometer
-several shear rates)

628

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Seiscor/Han Rheometer (on-line slit rheometer to measure


exit pressure)
AKSOOO Particle Sampling System
Shimadzu Scientific Instruments, Inc.
7102 Riverwood Drive
Columbia, Maryland 21046
(301) 381-1227
CFT-500 Flowtester (high-stress, weight-driven capillary
rheometer)
CFT-20 Flowtester (small, weight-driven capillary rheometer)
CMD Automatic Mooney Viscometer
Testing Machines Inc.
400 Bayview Ave.
Amityville, NY 11701
(516) 842-5400
Ray Ran Melt Flow Indexers
Mooney Viscometer
Time-Temperature Instruments Inc.
P. O. Box 40156
Pittsburgh, PA 15201-0156
(412) 621-5009
MECA Creep Rheometer (high-precision constant stress
rheometer for creep and recovery)
Tinius Olsen Testing Machine Company, Inc.
P. O. Box 429
Willow Grove, PA 19090-0429
(215) 675-7100
Extrusion Plastometer (melt indexer)
Sieglaff-McKelvey Rheometer (capillary rheometer)
Toyo Seiki Seisaku-Sho, Ltd.
15-4, 5-Chome
Takinogawa, Kita-Ku
Tokyo 114, Japan

APPENDIX E

629

Westover Rheometer (very high pressure capillary rheometer)


Capirograph (capillary rheometer)
Labo-Plastomil (torque rheometer)
Melt Indexer
Zwick of America, Inc.

P. O. Box 997
East Windsor, CT 06088
(203) 623-9475
Extrusion Plastometers

Nomenclature
NOTE

Equation number refers either to the definition of the quantity or


to the principal equation in which the quantity appears. Some
symbols that are used only once and defined in the text are not
listed here.
ROMAN LETTERS

a Various empirical constants, or length characteristic of


aT

A
AG
Ai
b
B

Boo
BA
BD
BH

Bw
Bij
c
C
Co
Cij
d
df
630

chemical structure of molecule (2-95)


Shift factor for time-temperature superposition
Surface area or interfacial area, or strain scale factor in
exponential shear (5-75), or empirical constant (15-3)
Linear viscoelastic property defined in (2-74) and (2-75)
Amplitude of component Ei of the electric vector
Various empirical constants or Rabinowitch correction
(8-20)
Extrudate swell ratio for capillary (8-62) or empirical constant (15-2, 15-3)
Ultimate swell ratio for capillary extrudate
Area swell ratio of parison (16-4)
Diameter swell ratio of parison (16-1)
Thickness swell ratio of parison (16-2)
Weight swell ratio of parison (16-3)
Component of Finger tensor
Instrument compliance or velocity of light
Stress optical coefficient or constant in Equation 15-5
Constants in WLF equation (2-129)
Component of Cauchy tensor
Tube diameter in Doi-Edwards model
Flight clearance

NOMENCLATURE

631

D Diameter or drop shape parameter (11-6)

Db Diameter of barrel
D Diameter of the root of the screw
Do Diameter of capillary die
D(t) Tensile creep compliance
D /Dt Substantial derivative 0-59)
e Bagley end correction (8-48) or flight width in extruder
04-4)
e i Unit normal vector (i = 1, 2 or 3)
E Young's modulus 0-12)
E Extruder power input 04-7)
E f Contribution of clearance flow to extruder power 04-14)
Ea Activation energy for flow (2-128) and 00-15)
Ei Component of electric field vector
E(t) Tensile relaxation modulus
flEoo Constant in Equation 12-5
fa Fractional free volume
F Force
F(A) Relaxation spectrum function (2-28)
Fij Component of displacement gradient tensor (A-4)
g Acceleration due to gravity
gi Relaxation strength of ith element of model (5-41, 5-44)
g/i) Relaxation strength for tensile recoil (6-22)
G Shear modulus 0-13)
G i Modulus of ith Maxwell element
G c Crossover modulus (2-78)
G d Amplitude ratio (lTo/Yo) in oscillatory shear
Gr Reduced modulus (2-119)
G(t) Shear stress relaxation modulus
G~ Plateau modulus
G'(w) Storage modulus
G"( w) Loss modulus
G*(w) Complex modulus
G~ Apparent value of storage modulus including the effect of
instrument compliance
G:; Apparent value of loss modulus including the effect of
instrument compliance
h Plate spacing or rheometer gap
hi Damping function of ith model element (5-27)
h( ) Damping function

632

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

HOt)
H
H(Il' 12 )
II
12
13

J
Jr
J(t)

JJ

J'(w)
J"( w)
J*(w)

k
K

Relaxation spectrum function (2-29)


Channel depth
Damping functional (3-82)
First scalar invariant of a Cartesian tensor
Second scalar invariant of a Cartesian tensor
Third scalar invariant of a Cartesian tensor
Shear compliance
Reduced compliance
Shear creep compliance
Steady state compliance in limit of very small shear stress
Storage compliance
Loss compliance
Complex compliance
Thermal conductivity of melt
Empirical constant or rheometer geometry constant (7-1,
7-2)

Ke Spring constant
Kv Dash pot viscous resistance coefficient
K", Constant in Equation 12-5
L Length of sample or of capillary or slit or of extruder
barrel
L Tube length in Doi-Edwards model
Lo Initial length of sample
m Various empirical constants or parameter for multiaxial
extensional flows (6-38)
met) Memory function (3-17)
M Molecular weight of monodisperse polymer or torque
Mi Molecular weight of component i
M( ) Strain-dependent memory function (3-61)
Me Critical molecular weight for appearance of entanglement
effects in the dependence of 7]0 on molecular weight
M: Critical molecular weight at which JJ ceases to depend on
M

Mi Molecular weight of material having weight fraction -w:


Mo Monomer molecular weight or torque amplitude in oscillatory shear for rotational rheometer or parameter of lognormal molecular weight distribution (B-6)
Me Average molecular weight between entanglements
Mn Number average molecular weight (see Appendix B)

NOMENCLATURE

633

Mw Weight average molecular weight (see Appendix B)


M z Z-average molecular weight (see Appendix B)
MI Melt index
n Various empirical constants or index of refraction
n Outer-directed unit normal vector
n ij Birefringence tensor
n' Real component of complex refractive index
nil Imaginary component of complex refractive index
N Angular screw speed
No Avogadro's number
N I ( y) First normal stress difference (simple shear)
Ni y) Second normal stress difference (simple shear)
Nit) First normal stress difference during transient shear deformation
Nit) Second normal stress difference during transient shear
deformation
N I ( t) + First normal stress growth function
Nit)+ Second normal stress growth function
p Ratio of disperse phase viscosity to matrix viscosity
P Pressure, or degree of polymerization, or energy dissipation per cycle per unit volume in oscillatory shear
Pa Ambient pressure
Pd Driving pressure-capillary flow
Pe Exit pressure
P* Hole pressure (8-36)
Po Extruder feed pressure
dP Pressure difference (P 2 - PI)
dPcap Pressure drop for fully-developed capillary flow
dPends Entrance loss plus exit loss (8-47)
dPent Entrance pressure drop
d Pex Exit pressure drop
PI Polydispersity index
Q Volumetric flow rate
Qd Drag flow rate
r Radial spatial coordinate in cylindrical coordinates
R Radius of rheometer fixtures or of blown film bubble, or
ratio of transmitted light intensity to incident intensity
2
<R ) Mean square distance between chain ends (10-1)
Rb Radius of barrel

634

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

Roo Ultimate recoil coefficient


R(t) Recoil function (recoverable compliance) (2-39)

Rg
Re
s
S

Soo
Sij
t
te
tE
fl.t
ti

to
tr
T

Ta
Tg
Tm
To

u
ui
U
v

VI
Vi

vbx
Vbz

vM
Vx

vz

Radius of gyration of molecule


Reynolds number (1-62)
(t - t') or interfacial tension
Linear strain or strength of network at the gel point (12-8)
or order parameter for liquid crystal polymer (17-1)
Ultimate tensile recoil function, Eoo/aE (see Equation 6-19)
(Equal to D2 for linear viscoelastic behavior)
Nonlinear strain tensor (3-65)
Time
Reentanglement time during interrupted shear (5-45)
Reentanglement time during reduction in shear rate (5-46)
Rise time in "step" strain experiment
Component of surface stress vector
A specific time at which strain is introduced into sample
Reduced time (2-120) or rest time during interrupted
shear
Temperature
Empirical constant in (15-6)
Glass transition temperature (see Appendix D)
Melting temperature
Reference temperature or wall temperature (7-4) or constant in (15-2)
Time-dependent elastic energy potential function (3-57,
3-58)
Displacement vector component in Xi direction
Strain-dependent elastic energy potential function (3-59)
Fluid velocity or velocity of light
Free volume
Velocity component in Xi direction
Transverse component of relative boundary plane velocity
Down-channel component of relative boundary plane velocity
Melt velocity in machine direction (blown film)
Transverse component of melt velocity in screw channel
Down-channel component of melt velocity in screw channel

NOMENCLATURE

635

V Velocity of moving plate or dimensionless down-channel


velocity
Vb Linear velocity of extruder barrel
~ Slip velocity (8-26)
Wi Weight fraction of component i
W Width of slit or perpendicular distance between flights
We Weber number (11-7)
W Work per cycle per unit volume of fluid
Xi Spatial coordinate or displacement function (A-I)
X M Machine direction coordinate (blown film)
x T Transverse direction coordinate (blown film)
X Reactive group conversion (12-1)
d X Plate displacement
y Radial coordinate in extruder equations
Y Dimensionless radial distance in extruder equations
z Distance of fluid element above a horizontal datum plane
or axial coordinate
GREEK LEITERS

a Empirical constant in (3-76), or exponential rate coeffi-

f3

'Y
'Yi
'Yo
'Yo
'Y,
'Y2

'Yoo

cient in exponential shear (5-75), or coefficient of thermal


expansion of fluid (D-l), or pressure coefficient of viscosity
(10-17), or thermal diffusivity of melt
Empirical constant in (3-79), or Rabinowitch correction for
slit flow (8-34), or angle between polarization direction
and principal direction of refractive index tensor, or modified Brinkman number (14-18), or parameter of log-normal
MWD (B-6)
Shear strain
Step strain introduced at time ti
Magnitude of shear strain at t = 0 in step-strain experiment
Strain amplitude in oscillatory shear
Recovered shear strain
Ultimate recoil (recoverable shear) for linear viscoelastic
behavior
Ultimate recoil (recoverable shear) (2-38)

636

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

'Y5 Shear strain at which

(J"

is maximum in stress growth

'YN Shear strain at which NI is maximum in stress growth

Y Shear rate

'Ym Mean shear rate in superposed steady and oscillatory

YA
YA 5

Yo

Yw

Y*

'Y.ij
'Yij

0i

0ij

{
C

c/t)
Coo

Co

iM
iT
iB
TJ
[ TJ]
TJi

TJ( y)

TJo

TJA
TJ!

TJ p
TJOi

TJoo
TJ +(t)

shear (5-69)
Apparent wall shear rate (8-13)
Apparent wall shear rate corrected for wall slip (8-26)
Empirical constant with units of shear rate in Equation
15-2
Absolute magnitude of wall shear rate
Representative shear rate (8-21)
Component of infinitesimal strain tensor
Component of rate of deformation tensor
Surface tension of melt
Mechanical loss angle
Phase angle of component Ei of the electric vector
Kronecker delta (3-7)
Monomeric friction coefficient
Hencky strain
Tensile recoil function
Ultimate tensile recoil (6-19)
Magnitude of Hencky strain imposed at t = 0 in extensional step strain
Hencky strain rate
Extensional strain rate in machine direction (blown film)
Extensional strain rate in transverse direction (blown film)
Biaxial extensional strain rate (6-24, 6-25)
Viscosity
Intrinsic viscosity (see Appendix C)
Viscosity of ith Maxwell element (Gjlt)
Viscosity function
Zero shear viscosity
Apparent viscosity (8-15)
Viscosity of suspending fluid
Plastic viscosity (1-24)
Zero shear viscosity of component i
Constant in Equation 12-5
Shear stress growth coefficient

NOMENCLATURE

17 -(t)
17;(0
17;(t)
17ii B)
17E( y)

7j(i)
17EC
17 ES
17'(W)
17"(W)
17*(W)
17 e
17 P

17,
()o

Ae
Ad
Ao
Al

AR
Ai
As
P
Po
(I'

(l'w
(l'B
(l'E
(I'm

(l'min
(1'0

(l'ij

637

Shear stress decay coefficient


Tensile stress growth coefficient
Biaxial stress growth coefficient
Biaxial extensional viscosity
Extensional viscosity (uniaxial extension)
Extensional viscosity (uniaxial and biaxial extension)
Apparent extensional viscosity determined in entrance flow
Apparent extensional viscosity determined in melt strength
test
Dynamic viscosity
In-phase component of complex viscosity
Complex viscosity
Exponential viscosity (5-80)
Principal exponential viscosity (5-83)
Relative viscosity (11-2) and (C-l)
Cone angle of cone-plate rheometer fixtures
Angle between molecular axis and the direction of a liquid
crystal polymer or helix angle of screw flights or angle
between tangent to bubble curve, R(z) curve, and vertical
(z) axis, i.e., arctan dR/dz
Relaxation time or other characteristic time of fluid or
wavelength of oscillation of electric vector
Equilibration time in Doi-Edwards model
Diffusion time in Doi-Edwards model
Characteristic time of a material (a material property)
Terminal (longest) relaxation time
Longest relaxation time in the Rouse model (2-98)
Relaxation time of ith Maxwell element
Shear wavelength (7-10)
Melt density
Melt density at the reference temperature
Shear stress in simple shear
Absolute magnitude of wall shear stress
Net stretching stress, biaxial extension
Principal stretching stress (1-34)
Maximum stress during interrupted shear experiment
Minimum stress during reduction of shear rate experiment
Stress amplitude in oscillatory shear or yield stress
Component of the stress tensor

638

MELT RHEOLOGY AND ITS ROLE IN PLASTICS PROCESSING

0" + (t) Shear stress growth function


O"(t, 'Y) Shear stress relaxation function

Component of extra stress tensor


c/> Volume fraction of filler
c/>m Maximum packing fraction of filler
C/>O Angular amplitude for oscillatory shear in a rotational
rheometer
<I> Dimensionless pressure in extruder
X Extinction angle
'l'1( y) First normal stress coefficient
'l'i y) Second normal stress coefficient
"1'1,0 Limiting value of '1'1 as y ~ 0
'1'2,0 Limiting value of '1'2 as y ~ 0
"I't(t) First normal stress growth coefficient
"I';(t) Second normal stress growth coefficient
'1'1-(1) First normal stress decay coefficient
'1'2-(1) Second normal stress decay coefficient
w Frequency
n Rotational speed (rad/s)
'Tij

Author Index
Numbers shown in italics refer to reference lists at ends of chapters.
Abbott, T. N. G., 284, 295
Acierno, D., 290, 296
Adams, E. B., 137, 140, 152
Adams, N., 282, 295
Adolf, D., 420, 423
Agarwal, P. K., 208, 212, 213, 229
Agassant, J. F., 329, 343,499, 508, 538, 555,
595,600
Agrawal, A. R., 565, 566
Ajroldi, G., 519, 529
Akana, Y., 552, 556
Aldhouse, S. T. E., 319, 342
Allan, P. S., 498, 508
Allen, W. F., 553, 556
Amari, T., 411, 421
Anders, S., 362, 364
Anderson, R. D., 411, 413, 415, 416, 421
Andrade, E. N. de C, 290, 296
Andrews, R. D., 99, 102, 410, 421
Anturkar, N. R., 554, 556
Aris, R., 25, 41
Arman, J., 63, 69, 100, 101, 102
Armstrong, R. C, 27, 41, 85,101,107,151,
156, 161, 163, 176, 177, 346, 363
Asada, T., 359, 364, 431, 439
Ashare, E., 171, 172, 177,222,230,276,295
Astarita, G., 222, 230
Athey, A. J., 539, 555
Au-Yeung, V. S., 518,529
Awaya, H., 339, 344
Azzam, R. M. A., 352, 354, 363
Baba, S. M., 475, 489
Bagley, E. B., 320, 342
Baird, D. G., 311, 312, 315, 318, 324, 342,
343, 362, 364, 436, 437, 440, 495, 508
Bakerdjian, Z., 334, 343, 434, 440, 584, 600
Balin, P. L., 471, 489

Balke, S., 607, 612


Ballenger, T. F., 336, 338, 339, 344
Ballman, R. L., 249, 267,551, 556
Balta Calleja F. J., 498, 508
Bartels, CR., 166, 177
Bashara, N. M., 352, 354, 363
Basu, S., 520, 521, 529, 530
Bata, G. L., 407, 409
Batchelor, G. K., 275, 295
Baumgartel, B., 72, 101
Bayer, R. K., 247, 267,498, 508
Bennett, K. E., 249, 262, 267
Bentley, M. E., 517, 529
Berardinelli, F. M., 523, 530
Bergem, N., 338, 344
Bernstein, B., 127, 151
Berry, G. C, 78, 101, 164, 177, 208, 209,
229, 426, 439
Bessho, N., 190, 229
Bevis, M. J., 498, 508
Beyer, C E., 517, 529
Biesenberger, J. A., 415, 419, 422, 445, 484,
485,490
Billingham, N. C, 615, 616
Billmeyer, Jr., F. W., 613, 616
Binding, D. D., 324, 343
Binnington, R. J., 346, 363
Bird, R. B., 27, 41, 85, 101, 107, 151, 156,
157,161,176,225,230,313,314,342,360,
361, 364
Blacklock, J. E., 539, 555
Blake, J. W., 411, 413, 415, 416, 421
Blakeslee, III, T. R., 539, 555
Blow, M. M. J., 417, 423
Boger, D. V., 314, 342, 346, 362, 363
Bogue, D. C, 137, 140, 152, 331, 336, 338,
339, 343, 344, 359, 360, 363, 364
Boiko, B. B., 338, 344
Boles, R. L., 331, 343
639

640

AUTHOR INDEX

Booij, H. c., 60, 61, 66, 100, 130, 133, 137,


139,151,152,174,175,177,178,219,229
Booy, M. L., 453, 465, 466, 489
Borisenkova, E. K., 338, 344
Bormuth, H., 506, 508
Boukhili, R., 495, 507
Bowers III, G. H., 399, 409
Brenna, A., 282, 295
Brewer, G. W., 496, 508
Bright, P. F., 399, 409
Brizitsky, V. I., 332, 335, 343
Broadhead, T. 0., 562, 566
Brodkey, R. S., 274, 294
Brown, R. A., 346, 363
Bruker, I., 194, 229,281,283, 295
Bueche, F., 75, 101, 162, 177,378, 389
Burkert, S. J., 414, 415, 422
Burkhardt, U., 486, 487, 490
Busse, W. F., 247, 267, 399, 409
Butcher, A. F., 203, 204, 229
Cain, J. J., 550, 556
Cakmak, M., 523, 530
Calundann, G. W., 425, 439
Campbell, G. A., 544, 546, 547, 555
Cancio, L. V., 471, 489, 517, 529
Cantow, M. J. R., 374, 375, 389
Cao, B., 546, 555
Carley, J. F., 477, 489
Carr, S. H., 406, 409
Carreau, P. J., 162, 177
Carroll, D. R., 414, 422
Castro, J. M., 411, 416, 421
Cat ani, A. M., 407, 409
Chambon, F., 420, 423
Chan, D., 273, 274
Chan, H. L. W., 416, 422
Chan, Y., 400, 405, 409
Chapman, F. M., 397, 400, 408
Chapman, G. R., 539, 555
Charles, M., 320, 333, 343
Chatraei, Sh., 262, 267
Chattopadhyay, S., 159, 177, 328, 329, 343
Chella, R., 483, 490
Chen, I-Jen., 336, 338, 339, 344
Chen, K.-c., 332, 343
Chen, S. c., 493, 507
Cheng, D. C.-H., 390, 391, 408
Chiang, H. H., 163, 177, 505, 508
Chmiel, H., 304, 326, 341
Chou, C. H., 317, 342
Chow, A. W., 276, 295
Choy, I.-C., 415, 422
Christiano, J. P., 536, 555

Christiansen, E. B., l70, 177


Chu, E., 495, 499, 500, 507, 508
Chung, C. I., 477, 489
Chung, S. c.-K., 265, 268
Cieloszyk, G. S., 537, 555
Clark, E. S., 172, 177,400,409,497,508
Clegg, D. W., 270, 275, 294
Cline, A. W., 536, 555
Co, A., 554, 556
Cogswell, F. N., 242, 249, 258, 260, 266, 267,
322, 338, 343, 344, 386, 389, 401, 409,
435-438,440,515,516,518,520,526,529,
538, 555, 598, 600
Cohen, R. E., 163, 177
Colaluca, M. A., 414, 422
Cole, K. S., 90, 101
Cole, R. H., 90, 101
Collyer, A. A., 270, 275, 294
Connelly, R. W., 221, 230
Cook, K. S., 294, 297
Cooper, A. R., 607, 612
Cooper, S. L., 157, 176
Corbett, H. 0., 547, 555
Cox, R. G., 403, 404, 409
Cox, R. H., 174, 177
Cox, W. P., 173, 174, 177
Crady, D. L., 264, 268
Cragg, L. H., 613, 616
Crawley, R. L., 65, 100, 195-198, 229,282,
295, 334, 343
Crist, B., 166, 177
Crochet, M. J., 37, 41, 335, 344
Cross, M. M., 162, 177, 283, 295, 378, 389
Crothers, D. M., 614, 616
Crowder, J. W., 336, 338, 339, 344
Crowson, R. J., 399, 400, 409
Crozier, D., 419, 423
Currie, P. K., 149, 152
Curry, J., 565, 566
Curtiss, C. F., 27, 41, 85, 101, 107, 151
Czarnecki, L., 400, 409
D'yakov, K. D., 561, 565
Dannhauser, W. c., 90, 91, 93, 101
Darby, J. P., 510, 529
Darby, R., 290, 296
Darnell, W. H., 472, 489
Davies, J. M., 286, 296, 313, 342
Davis, H. L., 331, 343
Davison, S., 406, 409
deCindio, B., 290, 296
de la Lande, M., 499, 508
De Rossett, T. A., 334, 343
de Vargas, L., 312, 342

AUTHOR INDEX

Dealy, J. M., 150, 151, 152, 204-206, 219,


223-228, 229, 230,241,244,260,261,265,
266-268,270,285-287,290,292-294, 294,
296,297,326-329,334,343,414,422,511,
514,516-519,521, 529,554,555,557,562,
565, 566, 570, 571, 599
Dearborn, J. R., 519, 521, 522, 529
Degeneuve, G., 99, 508
deGennes, P. D., 81, 101,424,426,427,430,
439
DeKee, D., 176, 178
Delos, S., 417, 422
Demarmels, A, 151, 152,253,264-266, 267,
268
Demus, D., 424, 439
Denn, M. M., 174, 177, 307, 314, 336, 338,
339, 340, 342, 344, 550, 554, 556
Denson, CD., 261, 264, 267, 268
Derezinski, S. J., 493, 507
Dierckes, A C, 158, 177
Dijksman, J. F., 495, 507
Dillon, J. G., 614, 616
Dillon, J. H., 410, 421
Dillon, R. E., 339, 344
Dimitrov, M., 471, 489
Dobroth, T., 553, 556
Dodge, J. S., 285, 295
Doi, M., 79, 82, 83, 85, 101, 107, 146-148,
151,349,357,359,362,363,364,426,439
Donis, R., 399, 409
Dormier, E. J., 205, 229, 511, 529
Doshi, S. R., 150, 152, 225-228, 230
Dreiblatt, A, 565, 566
Drexler, L. H., 313, 319, 342,354,356,357,
363
Du, C-C, 538, 555
Dusi, M. R., 411, 413, 417, 419, 421-423
Dutta, A, 521, 529
Dynes, P. J., 420, 423

Edelman, R., 523, 530


Edwards, R., 554, 556
Edwards, S. F., 81-83, 101, 107, 146-148,
151,359,362,364,370, 389
Einaga, Y., 188, 229
Eise, K., 486, 487, 490
Eisenberg, A, 618, 621
EI Kissi, N., 318, 340, 342
Elbirli, B., 163, 177, 475, 489
Eley, R. R., 411, 421
Elia, A E., 498, 508
Elmendorp, J. J., 405, 406, 409
Enns, J., 411-413, 421
Ente, J. J. S. M., 552, 556

641

Ericksen, J. L., 426, 427, 439


Erwin, L., 483, 484, 490, 521, 530, 553, 554,
556
Everage, A E., 249, 267
Fan, X. J., 225, 230
Farber, R., 544, 555
Faucher, J. A, 294, 297
Fayt, R., 405, 409
Fellers, J. F., 359, 364
Fenner, R. T., 442, 489
Fernandez, D., 315, 342
Fernandez, F., 520, 521, 529, 530
Ferry, J. D., 33n, 41, 43, 50, 55, 64, 65, 75,
78,80,89,91,94, 100, 101, 219, 229,382,
389, 584, 600
Fetters, L. J., 166, 177, 357, 362, 364
Fifer, R. L., 523, 530
Figlan, J., 411, 421
Finger, F. L., 336, 344
Finlayson, B. A, 313-315, 342
Fisa, R., 495, 507
Fischer, E., 519, 529
Fleissner, M., 290, 296
Flory, P. J., 411, 421, 425, 439, 610, 612, 619,
621
Flumerfelt, R. W., 157, 176
Folkes, M. J., 399-401, 409
Folt, V. L., 414, 422
Fortuin, J. M. H., 279, 295
Fox, T. G., 78, 101, 164, 177,619,621
France, G. H., 327, 343
Franck, A, 258, 259, 267, 286, 296
Frattini, P. L., 357, 363
Freeman, W., 417, 422
Friedman, E. M., 593, 600
Friedrich, C, 175, 178
Fritch, L. W., 505, 508
Fritz, H. G., 510, 529, 560, 565, 565, 566
Fritzen, J. S., 411, 413, 417, 421
Froelich, D., 245, 267
Fujiki, T., 388, 389, 540, 555
Fujiyama, M., 339, 344
Fukada, M., 186, 188, 189, 196, 199, 228, 229
Fuller, G. G., 276, 295, 345-347, 360, 362,
363,364
Furches, B. J., 506, 508
Furumiya, A, 552, 556
Ganani, E., 283, 295
Garbella, R. W., 283, 295
Garcia, M. N., 414, 422
Garcia-Rejon, A, 334, 343, 514, 516-518,
529

642

AUTHOR INDEX

Garritano, R., 415, 419, 422


Garverick, S., 417, 422
Gaskins, F. H., 323, 343
Gates, P. c., 538, 555
Gauvin, R., 495, 507
Gavin, P. T., 280, 295
Gavins, J., 332, 343
Geiger, K, 283, 295
Gendron, R., 174, 177
Gent, AN., 290, 296
George, H. H., 363, 364
Gergen, W. P., 406, 409
Ghijsels, A, 407, 409, 552, 556,584, 600
Giacomin, A J., 290, 292, 293, 296, 297,562,
566
Gibson, A. G., 331, 343
Gillham, J. K, 411-413, 416, 421, 422
Githuku, D., 248, 267
Glasscock, S. D., 334, 343
Gleissle, W., 175, 178, 280, 282, 294, 295
Glenn, Jr., W. B., 493, 507
Godfrey, J., 417, 422
Goettler, L. A, 399, 409
Gogos, C. G., 37, 41, 158, 177,203, 229,332,
343, 442, 488, 489, 493, 507
Goldfarb, I. J., 416, 422
Goldstein, c., 285, 295
Gonzalez, H., 521, 530
Gonzalez-Romero, V., 415, 416, 418-420,
422
Gotro, J. T., 411, 417, 419, 421, 422
Gotsis, A D., 318, 324, 342, 436, 437, 440
Gattfert, A, 559, 565
Gottgetreu, S. R., 475, 489
Gottlieb, M., 285, 295, 313, 314, 342
Gouz, J. J., 503, 508
Goyal, S. K, 499, 500, 508
Graessley, W. W., 65, 80, 84, 85, 88, 95,
100-102, 166, 177, 182, 187, 190, 192, 193,
195-199,201,202,228,229,282,295,334,
343, 369, 370, 373, 377-379, 386, 389
Grattom, R. F., 286, 296
Greener, J., 221, 230
Gregory, D. R., 384, 389
Greygang, G. G., 503, 508
Griffin, A X., 570, 599
Griffith, R. M., 335, 344
Gross, L. H., 285, 295
Gulrich, L. W., 425, 439
Gupta, R. K, 544, 555
Gupta, V. B., 482, 489
Hadad, D. W., 411, 413, 417, 421
Haessley, W. P., 522, 530
Haghtalab, A, 292, 297
Hagler, G. E., 336, 338, 339, 344

Hagnauer, G. L., 413, 421


Halasz, L., 563, 566
Hale, A, 414, 422
Han, C. D., 90, 101, 173, 177,313-316,319,
320,333,342,354,356,357,363,390,397,
400,402,408,409,411,415,419,421,423,
437, 438, 440, 549, 550, 556
Hannon, M. J., 411, 421, 518, 521, 525,529,
590-593, 600
Hansen, M. G., 281, 295
Hanson, D. E., 205, 229
Harban, A A, 571, 599
Harding, S. W., 162, 177, 378, 389
Harran, D., 415, 420, 422
Hashimoto, N., 488, 490
Hassager, 0., 27, 41, 85, 101, 107, 151, 156,
161, 176
Hatz, R., 424, 439
Hauson, E. E., 410, 421
Hayashi, T., 497, 508
Hedvig, P., 417, 422
Hegele, R., 471, 489
Heinle, P. J., 414, 421
Heinz, W., 561, 565
Helfand, 85, 101,225, 230
Henderson, AM., 335, 344, 514, 529
Henry, R. L., 290, 296
Henze, E. D., 516, 529
Herrington, F. J., 547, 555
Herrmann, H., 486, 487, 490, 565, 566
Hess, J. E., 414, 421
Heuser, G., 279, 295
Hieber, C. A, 163, 177, 505, 508
Higashitani, K, 311, 342
Hill, Y., 290, 296
Hilton, D. c., 264, 268
Hinch, E. J., 360, 364
Hinrichs, R. J., 411, 421
Hoff, M., 417, 422
Hoffman, D. M., 411, 415, 417, 421
Hoftyzer, P. J., 370, 383, 384, 389, 619, 621
Holmes, L. A, 171, 172, 177
Honerkamp, J., 72, 101
Hong, C.-N., 317, 332, 342, 343
Horie, K, 419, 423
Horrion, J., 577, 600
Hostettler, J., 283, 295
Hou, T. H., 419, 423
Hoverty, P., 417, 422
Hrymak, AN., 500, 508
Hsu, C. c., 158, 177, 493, 507
Hsu, T., 249, 262, 267
Huang, D. c., 335, 344
Huang, T., 544, 547, 555
Huilgol, R. R., 150, 152
Hull, AM., 505, 508, 587, 600

AUTHOR INDEX

Huppler, J. D., 171, 172, 177


Hiirlimann, H. P., 227, 230, 290, 291, 293,
296
Hushower, M. E., 419, 423
Hutton, J. F., 281, 295,313, 342
Ide, Y., 258, 259, 267, 435, 438, 440
Ingen Housz, J. F., 475, 489
Inomiya, K, 78, 101
Inoue, T., 97, 98, 102
Insarova, N. I., 338, 344
Isayev, A I., 222, 230, 332, 335, 343
Isherwood, D. P., 471, 489
Isono, Y., 219, 229
Iwama, M., 615, 616
Iyengar, V., 554, 556
Jabarin, S. A, 523, 530, 579, 600
Jackson, S., 565, 566
Jaffe, M., 425, 439
Jakopin, S., 486, 487, 490
Janeschitz-Kriegl, H., 140, 152, 262, 267,
292, 297, 358, 359, 362, 364, 495, 507
Janssen, L. P. B. M., 442, 487, 489
Jenkins, J. T., 265, 268
Jerome, R., 405, 409, 577, 600
Johnson, J. F., 374, 375, 389
Jones, D. N., 547, 551, 555
Jones, L. G., 274, 294
Jones, T. E. R., 286, 296
Jong, W. R., 505, 508
Jongschaap, R. J. J., 222, 230, 284, 295
Joseph, D. D., 276, 295
Joyner, R. S., 471, 489, 517, 529
Jud, K, 495, 507
Kachin, G. A, 506, 508
Kaiser, H., 519, 529
Kalika, D. S., 174, 177, 307, 338, 339, 340
Kahenbacher, E. J., 554, 556
Kalyon, D., 517, 518, 521, 529, 538, 555
Kamada, K, 615, 616
Kamal, M. R., 334, 343, 362, 364, 407, 409,
415,418,422,423,434,440,495,500, 507,
508, 517, 518, 521, 529, 584, 600
Kambe, H., 419, 423
Kanai, T., 547, 555
Kataoka, T., 393, 397, 400, 408, 409
Katsaros, J. D., 406, 409
Katsyutsevich, E. V., 222, 230
Kausch, H. H., 495, 507
Kaye, A., 127n, 283, 295
Kearsley, E. A, 127, 151
Keenan, J. D., 419, 423
Keentok, M., 281, 295
Kemblowski, Z., 410, 421

643

Kemp, R. A, 283, 295


Kepes, A, 561, 565
Kerker, M., 349, 363, 424, 439
Keunings, R., 37, 41
Khachatryan, G. M., 561, 565
Khan, S. A, 132, 152, 152, 390, 408
Khanna, Y. P., 166, 177
Khatta, R. K, 417, 422
Kietz, W., 497, 508
Kim, H. T., 510, 529
Kim, J. T., 471, 489
Kim, S.-G., 495, 508
Kimura, M., 576, 600
Kimura, S., 188, 190-193, 229, 290-292,296
King, R. G., 402, 409
Kiss, A D., 357, 362, 364
Kiss, G., 359, 364, 435, 436, 440
Kitagawa, K, 95-97, 102
Kitano, T., 393, 397, 408
Klein, I., 442, 447, 460, 467, 473, 475, 479,
489
Knapper, K M., 284, 295
Kojima, C. J., 419, 423
Kojima, T., 220, 230
Kojimoto, T., 190, 229
Kolinski, A, 86, 101
Komatsubara, T., 503, 508
Kondo, A, 318, 342
Koopmans, R. J., 336, 344, 515, 529
Kotaka, T., 219, 229
Kouba, K, 538, 555
Kozicki, W., 317, 342
Kral, V., 538, 555
Kranbuehl, D., 417, 422
Krause, E., 279, 295
Krauskoff, L. G., 324, 343
Krieger, I. M., 285, 295, 296, 397, 408
Kruder, G. A, 471, 489
Krul, N., 520, 529
Kumar, N. G., 160, 177
Kumar, R., 166, 177,415, 419, 422
Kuo, C C, 416, 422
Kurata, M., 57, 100, 183-188, 193, 195, 196,
199, 200, 219, 228, 229, 290, 291, 292,
296, 615, 616
Kurtz, S. J., 334, 343, 537, 539, 547, 551, 554,
555,556
Kuzuu, N. Y., 85, 101, 426, 439
Kwack, T. H., 549, 556
Kwolek, S. L., 425, 439
Kwon, T. H., 331, 343
Labana, S. S., 421, 423
Lafleur, P. G., 495, 507
LaLiberte, B. R., 413, 421
La Mantia, F. P., 290, 296

644

AUTHOR INDEX

Lambright, A. J., 399, 409


Lane, J. W., 417, 422
Larson R. G., 54, 100, 107, 149, 151, 132,
137,143,145,149-151,152,175,178,227,
230, 238, 265, 266, 268, 430, 439
Latrobe, A., 499, 508
Laudouard, A., 415, 420, 422
Laun, H. M., 71, 73, 92, 99, 100, 101, 102,
122,134,135,140,151,152,171,176,178,
183,188,190,198,202,208-210,213,
215-217,228,229,243,244,247-259,264,
265, 266, 267, 273, 289-291, 294, 296,
309,315,322-324,327,332,340,341,343,
549, 556, 598, 600
Laurence, R. L., 249, 262, 267
Lawler, J. V., 346,363
Lawton, E. L., 577, 600
Leaderman, H., 78, 101, 372, 389
Leal, L. G., 346, 360, 363, 364, 404, 409
Leblans, P., 65,100,133,137,139,152,174,
175, 177, 178,248, 267
Lee, B. L., 414, 422
Lee, C. Y.-c., 55, 100,416, 422
Lee, D-S., 419, 423
Lee, K. H., 274, 294
Lee, L. J., 335, 344
Lee, T. S., 397, 400, 408
Lee, W. A., 619, 621
Lee, W. K., 598, 600
Leib, R. I., 399, 409
Lem, K-W., 90,101,173, 177,411,415,421
Lenz, R. W., 576, 600
Leslie, F. M., 426, 427, 439
Lin, C. M., 554, 556
Lin, S. H., 158, 177
Lin, Y. H., 149, 152
Lindt, J. T., 473-475, 489
Link, G., 92-94, 102, 286, 296
Linster, J. J., 227, 230, 258, 259, 267
Lipshitz, S. D., 419, 423
Liu, T. Y., 290, 291, 296
Liu, T.-J., 317, 332, 342, 343
Lobe, V. M., 397, 408
Lockyer, M. A., 271, 294
Lodge, A. S., 115, 117, 119, 151, 154, 176,
224,230,281-283,295,310,312,313,342
Lofgren, E. A., 523, 530, 579, 600
Lopez Cabarcos, E., 498, 508
Lopulissa, J. S., 284, 295
Lord, H. A., 493, 507
Lorntson, J. M., 518, 529
Lucchesi, P. J., 554, 556
Luenberger, D. G., 480, 489
Luo, X-L., 189, 229
Lupton, J. M., 338, 339, 344
Lutz, R. G., 406, 409

Ma, C.-Y., 395, 397, 398, 400, 408


Maalcke, R. J., 405, 409
Macdonald, I. F., 222, 230, 276, 295
Mackay, M. E., 346, 362, 363
Mackley, M. R., 319, 342
MacKnight, W. J., 416, 422
Macosko, C. W., 71,101,140,152,249,262,
264, 267, 268, 281, 285, 295,411,413-416,
418-420,422,423,518,529,561,564,566
MacSporran, W. c., 283, 295
Maddock, B. H., 473, 477, 489
Malguarnera, S. c., 414, 422,495, 508
Malkin, A. Ya., 166, 169, 172, 177,419,423
Malone, M. F., 406, 409
Mamada, A., 415,422
Manaresi, P., 615, 616
Manisoli, A., 495, 508
Manzione, L. T., 411, 414, 421, 422
Marco, c., 577, 600
Marianucci, E., 615, 616
Marin, G., 63, 69, 95,100, 101, 102
Maron, S. H., 393, 408
Marrucci, G., 432, 439
Marsh, B. C., 276, 295
Marsh, B. D., 222, 230
Martin, G. c., 411, 419, 421
Martin, J. E., 420,423
Maskell, S. G., 515, 526, 529
Mason, S. G., 403, 404, 409
Masuda, T., 95-98, 102, 220, 223, 230, 285,
296, 552, 556
Matsumoto, T., 223, 230, 285, 296, 359, 360,
363,364
Mavridis, H., 500, 508
Maximovich, M. G., 411, 413, 417, 421
Maxwell, B., 205, 215, 229, 284, 286, 294,
295-297, 511, 529
May, C. A., 410, 411, 413, 417, 419, 421, 423
Mazich, K. A., 406, 409
McCarthy, R. V., 294, 297
McChesney, C. E., 425, 438, 439, 440
McGlamery, R. M., 571, 599
McIntyre, L. V., 225, 230
McKelvey, J. M., 450, 469, 477, 481, 482, 489
McKenna, G. B., 219, 229, 282, 295
McLeish, T. C. B., 85, 101
McNally, D., 438, 440, 495, 507
Mead, D. W., 290, 291, 296
Meijer, H. E. H., 475, 489
Meins, W., 498, 508
Meissner, J., 99, 102, 119, 134, 138, 140, 142,
151, 152, 161, 177, 181, 198, 199, 208,
212-215,227,228-230,243,244,247,249,
250,253,256-258,262,264-266,266-268,
273,282,283,289-291,293,294-296,548,
556

AUTHOR INDEX

Mendelson, R. A., 336, 344


Menezes, E. V., 199, 201, 202
Merz, E. H., 173, 174, 177
Metzner, A. 8., 390, 394, 400, 408, 426, 439,
544,555
Michaeli, W., 332, 343
Michele, J., 399, 409
Middleman, S., 37, 41, 158, 176,327, 343,
447, 467, 469, 489
Mihara, S., 540, 555
Mikkelsen, K. J., 357, 364
Mikols, W. J., 416, 422
Miller, B., 317, 342, 394, 408
Miller, J. c., 518, 529, 537, 555
Min, K., 390, 408, 395, 397, 398, 400, 408
Minagawa, N., 395-397, 400, 408
Minnick, L. A., 174, 177
Minoshima, W., 386, 389,550, 556
Mita, I., 419, 423
Mitsoulis, E., 314, 342
Mittal, R. K., 482, 489
Mochimaru, Y., 276, 295
Modan, M., 332, 343
Mokhtarian, F., 484, 490
Mol, E. A. J., 472, 489
Mondvai, I., 563, 566
Monge, Ph., 63, 69, 100, 101, 102
Montfort, J. P., 63, 69, 100, 101, 102
Moon, T. J., 227, 230, 290, 291, 296
Mooney, M., 567, 599
Moore, I, P. T., 319, 342
Morgan, P. W., 425, 439
Morganelli, P., 420, 423
Morgon, B. T., 517, 529
Morikita, N., 497,508
Morris, V. L., 419, 423
Morse, D. J., 281, 295
Muller, R., 245, 267
Muller, S. J., 346, 363
Munari, A., 615, 616
Muni, K., 411, 421
Miinstedt, H., 242-244, 250-259, 266, 267,
540,555
Murakami, K., 188, 229, 410, 421
Muramatsu, H., 359, 364
Murayama, T., 290, 296
Mussatti, F. G., 418, 421, 423
Mutel, A., 390, 408
Myers, A. W., 294, 297
Nabata, Y., 415, 422
Nadkarni, V. M., 159, 177, 328, 329, 343
Nagatsuka, Y., 393, 408
Nakajima, A., 552, 556
Nakamichi, T., 411, 421
Nakamura, K., 208, 209, 229

645

Nam, S., 539, 555


Narain, A., 276, 295
Nazem, F., 281, 290, 295, 296
Nelson, B., 562, 566
Nettelnbreker, H.-J., 565, 566
Nichols, R. J., 487, 490
Nielsen, L. E., 390, 394, 400, 408
Nishijima, K., 393, 397, 400, 408, 409
Nishizawa, K., 183, 184, 195, 228
Obron, S. J., 70, 101
Oda, K., 172, 177,400,409,410,421,497,

508

Ohta, S., 189, 196, 199, 220, 229, 230


Olbricht, W. L., 404, 409
Onogi, S., 95-98, 102, 223, 230, 285, 296,
359, 364, 431, 439
Onsager, L., 425, 439
Onuki, A., 349, 363
Ophir, Z., 435, 438, 440
Orbey, N., 334, 343,515-519, 521, 529
Orchard, S. E., 411, 421
Osaki, K., 135, 152, 183, 184, 186-193, 195,
196, 199,219, 228, 229,290,291, 292, 296
Otsubo, Y., 411, 421
Ottino, J. M., 483, 490
Oyanagi, Y., 78, 101, 336, 344, 393, 397, 400,

405, 408, 409

Palmen, J. H. M., 61, 100, 130, 137, 151,


152, 174, 175, 177
Pan dalai, K., 274, 294
Pandelidis, I. 0., 565, 566
Papanastasiou, A. C., 71,101,140,152,262,
267
Pappas, L. G., 414, 422
Park, J. Y., 550, 556
Park, W. S., 65, 100, 195-198, 229
Parnaby, J., 510, 529
Patel, P. D., 68, 100
Patzold, R., 399, 409
Payvar, P., 284, 295
Pearce, E. M., 437, 438, 440
Pearce, P. J., 413, 421
Pearson, D. S., 85, 101,225, 230,280, 295,
357, 362, 364
Pearson, J. R. A., 330, 343
Pecht, M., 565, 566
Penwell, R. c., 327, 343
Perdikoulias, J., 538, 555
Perez, G., 388, 389
Perry, S. J., 411, 416, 421
Petrie, C. J. S., 137, 152,241, 266,336, 344,
550, 554, 556
Phan-Thien, N., 312, 342

646

AUTHOR INDEX

Philipon, S., 499, 508


Philippoff, W., 227, 230, 289, 296, 320, 323,
333,343
Piau, J. M., 318, 340, 342
Pierce, P. E., 393, 408
Pie rick, M. W., 157, 176
Pieries, R. N., 471, 489
Pike, R. D., 311, 312, 315, 342
Pilati, F., 615, 616
Pipkin, A. C., 220, 230, 311, 342
Pisipati, R., 495, 508
Plazek, D. J., 70, 92,101,102,208,212,213,
229, 286, 296, 415, 422
Plochocki, A. P., 406, 409
Podolsky, Y. Y., 332, 335, 343
Pohl, H. A., 203, 229
Pollack, A., 521, 530
Porter, R. S., 327, 343, 359, 364, 374, 375,
389, 576, 593, 600
Portman, P., 264, 265, 268
Powell, R. L., 273, 274, 283, 295
Prasadarao, M., 437, 438, 440
Prettyman, I. B., 410, 421
Prichard, J. H., 388, 389
Prilutski, G. M., 426, 439
Pritchard, W. G., 311, 342
Pritchatt, R. J., 510, 529
Proctor, B., 537, 538, 555
Prud'homme, R. K., 390, 408
Quinzani, L. M., 280, 295
Raadsen, J., 407, 409, 552, 556, 584, 600
Raghupathi, N., 70, 101
Raible, T., 253, 262, 265, 267
Rallison, J. M., 346, 363, 404, 409
Ramachandran, S., 170, 177
Ramamurthy, A. Y., 306, 338-340,341,344,
511, 529, 539, 555, 595, 600
Ramanathan, R., 324, 343
Ramirez, H., 223, 230
Rauwendaal, c., 315, 342, 441, 443-445, 460,
461,464,466,467,470,475,477,484,488,
489, 490, 520, 529, 538, 555
Read, M. D., 311, 312, 315, 342
Read, W. T., 290, 295
Reddy, J. N., 312, 342
Reddy, K. R., 313, 332, 335, 342, 343
Regester, R. W., 338, 339, 344
Revesz, H., 563, 566
Richter, L., 424, 439
Rhum, D., 411, 421
Rice, P. D. R., 515, 526, 529
Richardson, S. M., 503, 505, 508, 587, 600
Richter, E. B., 411, 421
Ricketoson, R. c., 505, 508

Riggs, J. P., 425, 439


Ritzau, G., 205, 229
Rivlin, R. S., 290, 296
Roberts, E. H., 554, 556
Rochefort, W. E., 225, 230, 280, 295
Roger, M. G., 336, 344
Rogers, Y. G., 55, 100
Rohn, C. L., 92, 102, 561, 564, 565
Rokudai, M., 388, 389, 540, 555
Roller, M. B., 411, 414, 419,421,423
Roman, J. F., 334, 343
Roovers, J., 85, 101
Rose, W., 495, 507
Rouse, Jr., P. E., 74, 101
Roxbury, M. L., 290, 296
Roylance, M. E., 413, 421
Rubin, I. I., 503, 508
Rudin, A., 205, 229, 335, 344, 511, 514, 529,
539, 555, 570, 578, 599
Rumscheidt, F. D., 403, 409
Rutherford, R. A., 619, 621
Ryan, M. E., 415, 422, 495, 507, 521, 522,
529
Ryskin, G., 360, 364
Saillard, P., 538, 555, 595, 600
Saini, D. R., 159, 177, 328, 329, 343
Sakai, T., 488, 490
Salee, G., 576, 600
Sampers, J., 133, 137, 139, 152, 175, 178,
248, 267
Samurkas, T., 151, 152, 227, 230, 265, 268,
293, 297, 562, 566
Sandford, c., 397, 400, 409
Sasahara, M., 393, 397, 408
Saunders, S. W., 290, 296
Sawan, S. P., 411, 421
Sawyers, K. N., 114, 151
Scarola, L. S., 537, 539, 555
Schaefgen, J. R., 425, 439
Schaul, J. S., 518, 521, 525, 529, 590-593,
600
Schmid, H., 362, 364
Schmidt, R. L., 346, 363
Schmitz, K. P., 538, 555
Schneider, N. S., 416, 422
Schowalter, W. B., 285, 295
Schreiber, H. P., 205, 229, 511, 529, 570,
578, 599
Schuch, H., 69, 101,247-249,258,259,264,
265,267,322-324,327,332,343,549,556,
598,600
Schuler, A. N., 576, 600
Schulken, R. M., 174, 177
Schiimmer, P., 304, 326, 341
Schwartz, W. H., 283, 295
Schwarzl, F. R., 92-94, 102, 286, 296

AUTHOR INDEX

Scriven, L. E., 71, 101, 140, 152 262 264


267,268
'
,
,
Sebastion, D. H., 519, 521, 522, 529
Secor, R B., 264, 268
Seferis, J. c., 416, 419, 422, 423, 498, 508
Segawa, Y., 223, 230, 285, 296
Selopranoto, J. H., 505, 508 587 600
Semjonow, V., 169, 177
'
,
Senich, G. A, 416, 422
Senturia, S., 417, 422
Serrano, S., 415, 422
Seth, B. J., 130, 151
Seto, S., 414, 422
Shah, B. H., 290, 296
Sharma, P. K, 482, 489
Shaw, M. T., 163, 177, 334, 343, 414, 422
Shen, E. F., 331, 343
Shenoy, A V., 159, 177, 328, 329, 343
Sheppard, N., 417, 422
Sheptak, N., 517, 529
Shetty, R., 550, 556
Shih, C. K, 407, 409
Shirodkar, P., 249, 262, 267
Shirota, T., 393, 408
Shroff, R N., 249, 267
Shumsky, V. F., 172, 177
Shurcliff, W. A, 346, 348-350 352 363
Sieglaff, C. L., 414, 421
"
S!mmons, J. M., 168, 177
SltZ, C. E., 414, 422
Siva shinsky, N., 227, 230, 290, 291, 296
Skolnick, J., 86, 101
Sm!th, F. P., 205, 229, 511, 529
Smith, R. G., 78, 101,372, 389
Somer, K, 362, 364
Soong, D. S., 227, 230, 290 291 294 296
297
'
,
,
,
Sorta, E., 290, 296
Soskey, P. R, 137, 152, 262 267 283 295
Sosulin, K N., 561, 565
'
,
,
Souffie, RD., 539, 555
Sp~ncer, R S., 339, 344
Spiers, R. P., 283, 295
Spruiell, J. E., 523, 530
Srinivasan, R, 313, 342
Starita, J. M., 415, 419, 422 561 564 565
566
'
,
,
,
Starr, F. c., 294, 297
Stastna, J., 176, 178
Stephenson, S. E., 141-144, 152, 193, 229,
262, 264, 265, 267, 268
Stevens, M. J., 441, 457-459, 477, 489
Stevenson, J. F., 265, 268, 335, 344
Steward, E. L., 536, 555
Stockmayer, W. H., 615, 616
Stoehrer, B., 565, 566
Stratton, R A, 163-165, 177,203,204,229

647

Strel'tsov, A A, 561, 565


Struglinski, M. J., 166, 177
Sugeng, F., 312, 342
Suh, N. P., 495, 508
Sundstrom, D. W., 414, 415, 422
Swerdlow, M., 520, 529
Tadkmor, Z., 37, 41,158,177,332,343,442,
447,460,467,473,475,479,488 489 493
495,507
'
,
,
Tajima, Y. A, 419, 423
Takatori, E., 186, 228
Takigawa, T., 220, 230
Tamura, M., 188, 219, 229
Tan, V., 362, 364,407,409,517,529
Tanaka, H., 397, 400, 408, 409
Tanner, R I., 107, 128, 150, 151, 152, 168,
172, 177, 189, 219, 225, 229, 230, 238,
266, 281, 284, 295, 311, 313 332 335
342-344
'
,
,
Taylor, C. A, 522, 530
Tee, T. T., 223, 224, 230, 285, 296
Teyssie, Ph., 405, 409, 577, 600
Thann, R c., 539, 555
Thomas, A, 286, 296
Thoone, J. H., 60, 66, 100
T?rasher, KG., 411, 413, 417, 421
Tlemersma-Thoone, G., 174, 175, 177
Tiu, c., 317, 342
Tob?lsky, A V., 99, 102, 188, 229,410, 421
Tokl, S., 290, 296
Tokita, N., 130, 139, 151
Tong, P. P., 205, 229,511, 529
Tonogai, S., 414, 422
Tordella, J. P., 362, 364
Torza, S., 403, 404, 409
Torzecki, J., 410, 421
Tremblay, B., 318, 319, 340, 342
Troup, G. J., 346, 363
Tsai, AT., 227, 230, 290, 291, 294, 296, 297
Tsang, W. K W., 204-206, 225, 229, 230,
511, 529
Tschoegl, N. W., 43, 100
Tsunashima, Y., 615, 616
Tuminello, W. H., 68, 69, 100, 574, 600
Tuna, N. Y., 313-315, 342
Tung, C. Y. M., 420, 423
Tungare, A V., 411, 419, 421
Turner, S., 401, 409
Ulmer, AS., 498, 508
Umbach, H., 498, 508
Unsworth, J., 416, 422
Ushida, Y., 552, 556
Utracki, L. A, 174, 177, 334, 343, 407, 409
434, 440, 584, 600
'

648

AUTHOR INDEX

Valamonte, D., 471, 489


Valesano, V. A., 143, 145, 152
Valles, E. M., 280, 295
Van Aken, J. A, 262, 267
van de Hulst, H. c., 349, 363
van der Veen, A, 565, 566
Van Der Vegt, A K, 405, 409
Van Krevelen, D. W., 358, 364, 370, 383,
384, 389, 619, 621
Van Dene, H. J., 405, 409
Van Vijngaarden, H., 495, 507
Vergnes, B., 595, 600
Villemaire, J. P., 329, 343, 499, 508
Vincent, M., 499, 508
Vinogradov, G. V., 166, 169, 172, 177,222,
230, 332, 335, 338, 343, 344, 397, 409
Viola, G. G., 437, 440
Vlachopoulos, J., 314, 342, 500, 508, 538,
555
Vlcek, J., 538, 555
Volgstadt, F. R., 414, 421
Vrentas, C. M., 182, 187, 190, 192, 193, 228
Vu, T. K P., 285, 296
Wagner, H. L., 376, 378, 384, 389, 614, 616
Wagner, M. H., 54,100,126,130,134,
138-144,151, 152, 193, 198,209,213,215,
229, 240, 249, 251, 256, 257, 266-268
Walters, K, 271, 294, 283, 284, 295, 303,
309,313,341, 342
Wang, K K, 331, 343,505, 508
Warashina, Y., 223, 230, 285, 296
Watanabe, K, 411, 421
Watanabe, R, 359, 364
Webb, P. c., 515, 526, 529
Weeks, J. c., 515, 526, 529
Weese, J., 72, 101
Weissberg, H. L., 321, 343
Weissert, F. c., 395, 397, 398, 400, 408
Wereta, Jr., A, 411, 413, 417, 421
Werner, H., 486, 487, 490
Wesselling, P., 495, 507
Westover, R F., 385, 389
White, J. L., 130, 139, 151, 172, 174, 177,
181, 228, 258, 259, 267, 290, 296, 318,
334-336,338,339,341,342-344,359,364,
386,389,390,395,397,398,400,405,408,
409, 497, 508, 523, 530, 547, 550, 555,
556, 579, 600
White, Jr., R P., 415, 422
White, S. A, 318, 324, 342, 362, 364
Whorlow, R W., 303, 341
Wiest, J. M., 360, 361, 364
Wiff, D. R., 55, 100
Wilcoxon, J. P., 420, 423
Willey, S. K, 498, 508
Williams, C. E., 577, 600

Williams, G., 219, 229, 493, 507


Williams, G. E., 438, 440
Williams, H. L., 416, 422
Williams, J. G., 495, 507
Williams, L. c., 78, 101, 372, 406, 389
Williams, M. c., 290, 291, 296
Wilson, G. F., 510, 529
Wilson, N. R., 517, 529
Winter, H. H., 72, 101, 137, 152, 181, 228,
247,249,262, 267, 283, 295,406,409,420,
423, 510, 529,549, 551, 556
Wissbrun, K F., 55, 100, 175, 178,247, 267,
359,360,364,373,376,378,379,384,389,
411,421,426,432-437,439,440,518,521,
523,525,529,530,544,555,570,578,579,
590-593,598, 599, 600, 609, 612
Wong, C. P., 208, 209, 229
Worm, AT., 539, 555
Wortberg, J., 538, 555
Worth, R. A, 510, 529
Worthoff, R. H., 304, 326, 341
Wrasidlo, W. J., 417, 422
Wu, R., 537, 555
Wu, S., 68, 69, 100, 404, 405, 409,573, 599
Wu, W. C. L., 516, 529
Yaloff, S. A, 417, 422
Yamada, N., 188, 229
Yamamoto, F., 438, 440
Yamane, H., 174, 177, 181, 228, 341, 344,
550, 556, 579, 600
Yamasaki, H., 415, 422
Yandrasits, M., 417, 422
Yang, B., 335, 344
Yang, M.-C., 261, 267
Yang, T., 285, 296
Yang, W. P., 411, 413, 415, 416, 421
Yanovsky, Yu., 222, 230
Yap, C. Y., 416, 422
Yaris, R., 86, 101,349,363
Yasuda, KY., 163, 177
Yeh, P., 349, 363
Yen, H.-C., 225, 230
Yokoi, H., 497,508
Yoo, H. J., 397, 400, 409
Yottsutsuji, A, 503, 508
Yu, J. S., 538, 555
Zachman, H. G., 498, 508
Zahorchak, A c., 578, 579, 600
Zapas, L. J., 127, 136, 140, 152, 182, 219,
228, 229, 282, 295
Zeichner, G. R, 68, 100, 561, 565
Zimm, B. H., 75, 101,614,616
Zukas, W. X., 416, 422
Ziille, B., 227, 230, 290, 291, 293, 296

Subject Index
Acceleration, 40-41
Acetal copolymer, 610
Activation energy for flow, 383, 584
effect of branching on, 388
Adhesives, 411
Aerodynamic forces on a film bubble, 541
Aerospace industry, 411
Air ring
design of, 546-547
role of in film blowing, 531
Alignment, of molecules, 150-151
American Society for Testing and Materials
(see ASTM)
Amplitude ratio, for oscillatory shear, 60
Anisotropy, 20
optical, 349
Annular flow, as viscometric flow, 156
Antithixotropy, 17
Apparent extensional viscosity
from converging flow test, 249
from melt spinning test, 248
Area swell (see Parison swell)
Arrhenius equation, 89, 383
Aspect ratio, of filler, 393
ASTM, (American Society for Testing and
Materials), 327, 328, 414, 415, SOl, 503,
596,614
Average molecular weights, 608-610
(see a.lso Number average; Weight average;
Z-average; Viscosity average)
Bagley end correction, 320-322, 325
effect of filler on, 400
Bannatek, 560
Barrier screws, 536
Bernoulli's equation, 544, 545
Biaxial extension, 234
experimental methods for, 262
role of in film blowing, 543
techniques for generating, 261
velocity distribution for, 260
Biaxial extensional viscosity, 261

Biaxial start-up flow, 261


Biaxial stress growth coefficient, 261
Bingham plastic, 18, 398
Birefrigence, 349
applications of, 362
form, 350
measurement of, 352-357
related to molecular orientation, 360
related to stress, 358
use of to observe entrance effects, 319
BKZ equation, 127, 129, 136, 149
separable form of, 128, 266
Blends
compliance of, 372
miscibility of, 406
Blends, immiscible
drop breakup in, 403
drop deformation in, 403
estimation of Tg for, 620
morphology of, 405
role of drop elasticity, 405
role of Weber number in, 403
size of disperse phase, 405
viscosity of, 401, 402
Block copolymers, 575
as immiscible blends, 402
rheological properties of, 407
Blow molding
description of process, 509
flow in the die, 510
of engineering resins, 522
resin evaluation for 524, 525
resin quality control for, 528
(see also Injection blow molding; Parison
swell; Parison sag; Pleating; Parison
formation; Parison inflation; Stretch
blow molding)
Blow up ratio (BUR), 531, 534
Blowability, 549
Blown film (see Film blowing)
Body force, 4, 5, 40
Bohlin Reologi, 622
649

650

SUBJECT INDEX

Boltzmann superposition principle, 44-47,


64,72-73, 104, 108, 112, 191, 193,208
and oscillatory shear, 60
and time-temperature superposition, 88
applied to creep compliance, 58
applied to recoil, 58
applied to step tensile strain, 46
generalization of for nonlinear
viscoelasticity, 115
limitations of, 106
Box function for relaxation spectrum, 99
Brabender Instruments (C. W. Brabender
Instruments), 622
Branching, long-chain, 51, 576, 585
and MWD, 386
effect of on activation energy, 388
effect of on linear behavior, 99
effect of on steady state compliance, 388
effect of on time-temperature
superposition, 92
effect of on viscosity, 386-388
Brinkman number, 467
Brownian motion, 19, 20, 42, 48-50, 75, 85,
357
and entanglements, 142
Bubble flow (see Film blowing)
Bubble stability, 535, 550
related to extension thickening, 237
Bueche-Ferry law, 79, 83
Bueche model (see Rouse model)
Bueche-Harding equation, 162, 378
Capillary flow
apparent shear rate, definition, 300
apparent viscosity in, 301
entrance pressure drop for, 322
of power law fluid, 301
Rabinowitch correction, 303
Schuiimmer approximation, 304
shear rate, Newtonian fluid, 299
shear stress distribution, 298
slip velocity, 306
wall shear rate, Newtonian fluid, 300
wall shear rate, power law fluid, 302
wall shear stress, 299, 321-322
wall slip in, 305
(see also Capillary rheometers; Channel
flow; Entrance effects)
Capillary rheometers, 324
apparent shear rate in, 326
effect of pressure on, 327
effect of wall slip, 327
for on-line use, 558
types of, 327
viscous heating in, 326
wall shear rate in, 325
wall shear stress in, 325

(see also Capillary flow; Entrance effects;


Entrance pressure drop; Bagley end
correction)
Carreau equation for viscosity, 162
Carri-Med, 622
Carter Baker Enterprises, 623
Casson equation, 399
Cauchy strain tensor, 111
Cauchy tensor, 109, 112
components for simple extension, 110
components for simple shear, 110
definition of, 605
(see also Cauchy strain tensor)
Cauchy's equation, 37, 39, 40
Ceast U. S. A., 623
Cessation of steady shear
for rubberlike liquid, 121
linear response, 73-74
nonlinear behavior, 199
Chain scission, 607
Channel flow
converging, 329
entrance effects, 317
irregular cross sections, 317
use of lubrication approximation, 329
Weissenberg number for, 331
(see also Converging flow)
Characterization of polymers, 573
Chemorheology, 410
Chromatography, 410
Coating (see sheet extrusion)
Coextrusion film blowing, 535
Coil-stretch transition, 360
Cole-Cole plot, 90
Collapsing ladder, 541
Complex compliance, 69
(see also Storage and loss compliances)
Complex modulus, 63
strain amplitude-dependent, 219
(see also Storage and loss moduli)
Complex viscosity, 64
Compliance, of polymers, (see Creep
compliance; Steady state compliance)
Compliance, of rheometers, 270, 271
correction for, 273
elimination of, 273
Compounding, use of on-line rheometers for,
565
Compressibility, role of in extrudate
distortion, 338
Compression, 5
Concentric cylinder rheometer, 285
Condensation polymers, 576, 579
Cone and plate flow, 277
as a viscometric flow, 157
(see also Rheometers)
Conformation, of molecule, 365

SUBJECT INDEX

Conservation of mass, 38
Constitutive equations, 2, 37, 52, 107, 108
Doi-Edwards, 148
for Newtonian fluid, 36
nonlinear, 107
(see also Continuum models)
Constrained recoil (see Recoil)
Continuity equation, 38, 39
Continuum, 38, 39
Continuum mechanics, 38, 107
Continuum models 107, 127
Contour length fluctuations, 85
Contour length relaxation, 147, 148
Convected Maxwell Model, 117
Converging flow, 236
Deborah number for, 330
lubricated, 249
pressure drop in, 330
tensile extension in, 323
use of lubrication approximation for, 329
use of to measure apparent extensional
viscosity, 249, 323
use of for quality control, 598
(see also Entrance effects; Entrance
pressure drop)
Copolymers, 576
estimation of Tg for, 620
Costech Associates, 623
Couette flow, as viscometric flow, 156
Coupling agents, effect of on suspension yield
stress, 397
Cox-Merz Rules, 173-175,584
for block copolymers, 407
Creep, 14,55
methods of measuring, 286
(see also Extensional creep)
Creep compliance, 55, 56
for typical polymer, 58, 59
nonlinear, 207
time-temperature superposition, 209
of polystyrene, 94
Creep recovery (see Recoil)
Creeping flow, 40, 41
Creepmeter
rotational, 286
sliding plate, 291
tensile, 242-243
Cross equation for viscosity, 162, 378
Cross-linking reactions, 410, 411, 607
gel point of, 420
of LLDPE for film blowing, 552
models for, 419
rate of, 419
relaxation modulus during, 420
storage and loss moduli during, 420
use of dielectric analysis to monitor, 417
use of rheometers to monitor, 413-414

651

Use of supported samples to study, 416


viscosity during, 418-419
Crystallinity, 595
Crystallization, 574-576
and glass transition, 619
Cuff effect (see Parison swell)
Cure time, 420
Curing reactions (see Cross-linking reactions)
Curtaining (see Pleating)
Curtiss-Bird Model, 85
Custom Scientific Instruments, 215, 623

Damping function, 131


and Cox-Merz rules, 174
and Gleissle mirror relationships, 175
and irreversibility, 142
and ultimate recoil following steady simple
shear, 216
comments on use of, 144
definition of, 130
dependence on temperature, 136
determination of, 131
equations for, 134
exponential, 135
for extensional flows, 133, 138, 239-240
for multiaxial flow, 266
for polystyrene, 184
for relaxation strengths, 190
from Doi-Edwards theory, 149
in terms of entanglements, 141
prediction of tensile and shear recoil using,
143
relationship to shear stress growth
coefficient, 196, 198
shear, 132
shear, equations for, 136
shear, for various polymers, 137
universal, 139
Damping functional, 142
Data acquisition systems, 572
Daventest Limited, 624
Deborah number
for converging flow, 330.
for die flow, 332
for oscillatory shear, 220
Deformation, 1, 6, 8, 605
stretching, 232
large, rapid, 103
Deformation gradient tensor, 35
Degradation, 607
controlled, of polypropylene, 565
correction of rheological data for, 580-581
due to moisture, 579
in film blowing, 536
use of rheology to monitor, 410

652

SUBJECT INDEX

Devolatilization, in extruders, 445, 484


Diameter swelHsee Parison swell)
Dichroism, 350
Die design, 332
Die flow
in blow molding process, 510
in coat hanger die, 595
Die swell (see Extrudate swell)
Dielectric analysis, 417
Differential scanning calorimetry, 410, 574,
618
Differential thermal analysis, 618
Disclinations, 430
Discrete spectrum, 53
Disengagement time, 83, 147
Displacement functions
definition of, 601
for simple extension,,606
for simple shear, 601
Displacement gradient tensor
definition of, 602
for simple .extension, 606
for simple shear, 605
Displacement vector, 32
Dissipation
and structural dependency, 16
effect of on temperature distribution in
rheometers, 274
in dash pot, 15
in extruders, 453, 461
in oscillatory shear, 63
Diverging flow, use of to generate biaxial
e~tension, 262
Doi-Edwards Constitutive Equation, 148
damping function from, 149
prediction of viscosity by, 149
Doi-Edwards theory, 82-85, 183, 184
damping function for shear, 137
double step strain, 192
nonlinear viscoelasticity, 146-149
prediction for double step strain, 193
strain dependent relaxation modulus
prediction, 147
stress growth coefficients for, 199
Doolittle equation, 382
Double-step reversing strain test
effect of irreversibility assumption, 143
Double step strain, 191
prediction of Wagner's equation, 192
Drag flow, 156
combined with pressure flow, 157
sliding surface, 269
Drag flow rheometers (see Rheometers)
Draping (see Pleating)
Draw down (see Parison sag)
Draw down ratio (DDR), 531, 534

Draw resonance
in sheet extrusion, 554
use of for quality control, 598
Drawability, related to extension thickening,
237
Ductile failure, in simple extension, 242
Dupont Company, 624
Dynamic mechanical analysis, used to study
curing reactions, 416
Dynamic spring analysis, 416
Dynamic viscosity, 63
Eccentric rotating disk rheometer (see
Rheometers)
Edge bead in sheet extrusion, 552
Edge effects (see End and edge effects)
Einstein summation convention, 25
Einstein's equation, 390
Elastic energy potential, 127
Elastic energy storage, 17
Elasticity, 10
Elastomer, 2
Electric vector, 347
Electronics industry, 411, 414
End and Edge Effects
in cone-plate rheometers, 280
in drag flow rheometers, 275
in rheometers, 158
End effects in capillary flow (see Entrance
effects; Entrance pressure drop; Exit
pressure; Bagley end correction)
End-to-end distance, of molecule, 19-20
Energy dissipation (see Dissipation)
Entanglement coupling, 79
nature of, 80
Entanglement density, 141
related to viscosity, 167
Entanglements, 21, 49-50, 80, 83, 358
effect of on viscosity, 164
evidence for existence of, 79
Entrance effects
for filled melts, 400
in channel flow, 317
observations of in capillaries and slits, 318
related to extrudate distortion, 339, 341
vortex flow pattern, 318
Entrance flow (see Entrance effects;
Entrance pressure drop)
Entrance pressure drop, 319
and viscoelasticity, 323
and extensional flow, 323
and extensional viscosity, 323
and recoverable shear, 323
for Newtonian fluid, 320-321
in melt indexers, 328

SUBJECT INDEX

related to orifice pressure drop, 322


Bagley end correction)
Epoxy curing, 415
Equation of state (see Constitutive equation)
Equibiaxial extension (see Biaxial extension)
Equilibration time, 83, 146
Equilibrium modulus of elastomer, 49
Exit pressure
determination of N\ from, 313-314
measurement of, 315, 560
Exponential shear, 225
compared with planar extension, 151, 265
Exponential viscosity, 226
Extensiometers (see Rheometers,
extensional)
Extension thickening behavior, 254
definition of, 235
effect of branching on, 260
ofHDPE,259
Extension thinning behavior
definition of, 235
Extensional creep, 58
(see also Tensile creep compliance)
Extensional flow, 231, 236
at entrance to capillary, 323-324
coordinate system for, 233
definition of, 232
genera], 265
multiaxial, 265
rate of deformation tensor for, 265
role of in blow molding, 527
Extensional recoil (see Tensile recoil)
Extensional rheometers (see Rheometers)
Extensional stress growth coefficient, role of
in film blowing, 543
Extensional viscosity (see Tensile viscosity)
Extinction angle, 360
Extinction coefficient, 349
Extinction of light, 354
Extra stress, 31
in Newtonion fluid in simple shear, 36
Extrudate distortion
critical stress for, 539
effect of die material, 340
effect of resin additives, 340
factors it depends on, 337
gross melt fracture, 339, 341
in blow molding process, 511, 527
in film blowing, 534, 538, 539
of HDPE, 338, 339
of LDPE, 341
of LLDPE, 339
of PDMS, 340
of polystyrene, 339
oscillatory flow, 338
related to entrance flow, 339
(see also

653

role of fluorocarbons, 539


role of slip, 340
spurt effect, 338
types of, 336
Extrudate drawing
analysis of, 247
to generate uniaxial extension, 246
use of to measure apparent extensional
viscosity, 247
isothermal, 248
(see also Melt spinning)
Extrudate swell
definition of, 332
dependence on temperature, 335
dependence on MWD, 335
dependence on time and temperature, 334
effect of filler on, 400
effect of LID on, 333
for capillary, 332
for Newtonian fluid, 332
for noncircular dies, 335
in film blowing, 540, 544
methods for measurement of, 334
. of block copolymer, 408
of HDPE, 333, 334, 336
of polypropylene, 334
prediction of, 335
related to N\, 335
theoretical analyses of, 335
ultimate value of, 334
use of for quality control, 597
(see also Parison swell)
Extruders, 441
adiabatic flow in, 469
analysis of, 441
characteristic curves for, 463-466
coupled with dies, 454
devolatilization in, 484
drag flow and pressure flow in, 451
effects of simplifying assumptions on, 459
efficiency of, 453
feed zone in, 470
functions of, 442
leakage flow in, 453, 460, 461
melt conveying zone, 446
melting zone of, 472
mixing in, 480-484
modelling of, 459-469
non-isothermal flow in, 467
operating diagrams for, 455-459
plasticating, 444
power consumption in, 461
role of Brinkman number in, 467
scale-up of, 476-480, 587
screw zones of, 444
single-screw, simple model of, 446-454

654

SUBJECT INDEX

Extruders (cont'd.)
solids conveying in, 470
twin screw, 485-489
types of, 443
use of power law to model, 468
velocity distribution in, 449-452
viscous dissipation in, 453, 461
Extrusion (see Extruders)
Extrusion blow molding (see Blow molding)
Extrusion casting (see Sheet extrusion)
Extrusion coating (see Sheet extrusion)
Extrusion plastometer (see Melt index)
Fading memory, 22, 47
Fiber-filled melts, flow of, 399
Filled melts
elasticity of, 400
entrance effects for, 400
extrudate swell of, 400
first normal stress difference of, 400
viscosity of, 393
(see also Viscosity)
Film blowing, 247
aerodynamic forces, 544, 546
air ring design, 546-547, 552
bubble shape, 542, 543, 547-548
bubble stability, 548, 550
bubble temperature, 547
cooling air flow, 545
cooling rate, 547
description of process, 531
die flow, 538
drawability, 549
extrudate distortion, 538, 539
extrudate swell, 536
flow in extruder and die, 536
flow in the bubble, 540
forces acting on bubble, 541-543
gauge variations, 551
linear and branched polyethylenes, 552
maximum draw ratio, 549
molecular orientation, 534, 544
objectives of process, 533
of polyethylene, 551
process optimization, 590
production problems, 534
resins used for, 536
role of branching, 549
role of MWD, 549, 551
role of rheological properties, 533
selection of resins for, 590
sharkskin, 533, 534
special extruders for, 536
strain history, 544
strain rates in bubble, 543, 544

use of blends for, 552


viscous stresses, 543
(see also Draw down ratio; Air ring;
Bubble stability)
Finger strain tensor, 111
Finger tensor, 109, 112, 130
components for simple extension, 110
components for simple shear, 110
definition of, 113, 605
for simple shear, 113
(see also Finger strain tensor)
Finite linear viscoelasticity, 108, 115
First normal stress coefficient
definition of, 155
dependence on relaxation spectrum, 171
effect of molecular weight on, 170
effect of shear rate on, 173
for rubberlike liquid, 120
related to storage and loss moduli, 176
related to viscosity, 176
relationship to steady state compliance,
121
relationship to storage modulus, 121
relationship to relaxation modulus, 120
temperature dependence of, 171
time-temperature superposition of, 173
First normal stress decay coefficient, 200
First normal stress decay function, 200
First normal stress difference, 170
for a purely elastic linear rubber, 123
for lart~e amplitude oscillatory shear, 224
of filled melt, 400
of LDPE, 316
of liquid crystal polymers, 437
of rubberlike liquid, 120
related to exit pressure, 313-314
related to shear stress, 172
relationship with steady state compliance
at low .y, 170
sources of error in measurement of, 282
use of cone-plate rheometer to measure,
281
use of hole pressure to determine, 312
use of slit rheometer to determine, 309
(see also First normal stress coefficient;
Normal stress differences)
First normal stress growth function, 195
for rubberlike liquid, 120
Flow in channels (see Channel flow)
Flow rate (see Melt index)
Fluid mechanics, 37
Fountain effect, 362, 495
Fourier series, for large amplitude oscillatory
shear, 223
Free volume, 382-383
role of in glass transition, 618-619

SUBJECT INDEX

Friction coefficient, of molecule, 75-76


Frost line, in film blowing, 531, 544
Gate flow (see Injection molding)
Gaussian normal error curve, 611
Gel point
definition of, 420
determination of, 420
Gel state, 411
Gel time, 420
Gelation (see Cross-linking reactions)
Generalized Maxwell model, 52-53
and Lodge's network theory, 117
and steady state compliance, 56
and time-temperature superposition, 87
determination of parameters for, 71
storage and loss moduli for, 64
Glass transition temperature, 19,50-51,382,
574
definition of, 617
effect of molecular structure on, 619
effect of plasticizers on, 619
measurement of, 617
of blends, 620
of copolymers, 620
related to melting point, 620
Glassy behavior
of polymers, 50-51, 66
of elastomer, 48
of melt, 49
Gleissle Mirror relations, evaluation of, 175
G6ttfert, 245, 559, 624
Rheotens, 247
UBI/Haake, 625
HDPE (High density polyethylene)
damping function, extension, 139
entrance flow of, 318
entrance pressure drop of, 320
extensional flow of, 258
extrudate distortion of, 338, 339
extrudate swell of, 333, 334, 336
film blowing of, 536, 548, 549, 551
MWD of, 611
parison swell of, 514
tensile viscosity of, 258
use of for film blowing, 535
Helical flow, as viscometric flow, 157
Hencky strain, 8, 35, 606
Hencky strain rate, 9, 237
High density polyethylene (see HDPE)
Hole pressure
definition of, 310
of Newtonian fluid, 310
on-line measurement of, 560

655

related to N!, 310-311


use of to determine N!, 312, 560
Hooke's Law
for extension, 10
for simple shear, 11
Hydrodynamic interaction, in polymer
solutions, 75
Hydroxypropylcellulose, 425
Immiscible blends, rheological properties of,
406
In-line rheometers, 562
compared with on-line rheometers, 558
(see also On-line rheometers)
Incompressibility of polymers, 29, 30, 38,113
Independent alignment assumption, 148, 149,
193
Infinitesimal strain tensor (see Strain tensor)
Infrared spectroscopy, 411
Inherent melt viscosity, 614
Inherent viscosity, 614
role of in injection blow molding, 523
Injection blow molding
description of process, 509
of PET, 523
stretch blow molding, 523
Injection molding
controlled rheology resins for, 506
description of process, 491
drooling, 499
elongational flow, 498
evaluation of resins for, 500
flash, 499
flow in runners, 492
fountain effect, 495
frozen wall layer, 495
gate flow, 494
microstructure formation, 500
modelling of, 499
mold cavity flow, 494
mold design for, 499
mold ability tests, 502
molecular orientation, 497
multiple live-feed, 498
of liquid crystal polymers, 438
of thin parts, 499, 506
optimal operating conditions, 498
process objectives, 491
residual stresses, 497
resin evaluation for, 505-506
rheological models for, 587
role of constitutive equation, 500
role of melt index, 501
role of MWD, 500-501
role of rheological properties, 502
selection of resins for, 506

656

SUBJECT INDEX

Injection molding (cont'd.)


shear rate in mold, 495
short shots, 494
use of birefringence to measure stresses,
362

use of on-line rheometers for, 565


viscosity models for, 499
(see also Runners; Weld line; Jetting;
Spiral mold test)
Instron, 625
Instrument compliance (see Compliance, of
rheometers)
Interfacial tension
effect of on blend viscosity, 402
Internal bubble cooling, 534
Interrupted shear, 203
Intrinsic viscosity, 615
Invariants (see Scalar invariants)
Inverse of a tensor, 603n
Irreversibility assumption, 144, 194
and Wagner's equation, 192
effect of on tensile and shear recoil
predictions, 143
in Wagner's equation, 142
Isotropic stress, 29, 30
IUPAC A (LDPE), 249-251, 257
IUPAC C (LDPE), 540
IUPAC X (LDPE), 227
Jetting, in injection molding
causes of, 497
elimination of, 497
Jones matrix, 352, 354
Karl Frank, 624
Kayeness, 625
Kelvin body (see Voigt body)
Killion, 625
Kinematics, 45, 103, 104, 150, 180, 232, 238,
260

change of with time, 151


Knit line (see Weld line)
Kronecker delta, 111
Large amplitude oscillatory shear (see
Oscillatory shear)
Laser Doppler velocimetry, 346
Laser speckle interferometry, 346
Layflat width, 532
LDPE (Low density polyethylene)
cessation of steady simple shear, 203
interrupted shear, 204
damping function, shear, 135, 198
damping function, extension, 138, 139, 257

edge fracture in, 281


entrance flow of, 318
exit pressure of, 315
exponential shear, 227
extensional flow properties, 253
extrudate distortion of, 341
film blowing of, 536, 539, 540, 547, 549,
551

film resins, 181


first normal stress difference, 315
melt strength of, 248
nonlinear creep, 213
nonlinear relaxation modulus of, 133, 188
planar extension of, 265
recoil during tensile start-up flow, 143
recoil following steady shear, 214, 216
recoil function of, 212
relaxation modulus of, 99
sheet extrusion of, 553
simple extension of, 244
start-up of steady simple shear, 134
steady state compliance of, 254
stress growth coefficient for, 198
stress growth functions, 214
tensile creep of, 243
tensile recoil of, 249
tensile stress growth function of, 126
tensile stress growth coefficient of, 250
tensile viscosity of, 254, 255
time-temperature superposition applied
to, 188
use of for film blowing, 535
viscosity of at several temperatures, 160
Leslie-Erickson constitutive equation,
427-428

Leslie-Erickson theory for liquid crystals,


426-427

Light absorption, 347-349


Light scattering, 347-349, 360
Light transmission, 354
Linear low density polyethylene (see
LLDPE)
Linear polymers, simple extension of, 258
Linear strain, for simple extension, 7, 35
Liquid crystal polymers (LCPs)
birefringence of, 359
commercially available, 425
domain model of, 432-433
extrudate swell of, 438
first normal stress difference of, 437
lyotropic, 424
molding of, 438
nematic, 424
processing of, 437
shape of molecule, 425

SUBJECTr INDEX

thermotropic, 424
validity of stress-optical relation for, 360
viscosity of, 431-437
Liquid crystals
continuum theory of, 426
disclinations in, 430
viscosity of, from L-E theory, 428
viscosity of, measured, 429-430
LLDPE (Linear low density polyethylene)
double step strain, 143
draw resonance in, 554
extrudate distortion of, 339
film blowing of, 536, 538, 539, 547, 549,
551, 552
sheet extrusion of, 554
slip velocity of, 306, 307
use of for film blowing, 535
Lodge-Meissner relationship, 133, 148, 190
Lodge Stressmeter, 560
Lodge's network theory, 117
use of to predict parison sag, 520
Log normal MWD, 610
Long-chain branching (see Branching)
Loss modulus (see Storage and loss moduli)
Low density polyethylene (see LDPE)
Lubricated squeezing flow, to generate
biaxial extension, 262
Lubricating oil, 2
Lubrication approximation, 329
for converging flow, 331
in simulation of melt processing, 330
Manufacturers of rheometers, 622
Mark-Houwink-Sakurada equation, 615
Maron-Pierce equation, 395, 615
Master curve, 87
Master Unit Die Products, 504
Material constant, 3, 11
Material functions, 3, 74
for steady simple shear, 155
of nonlinear viscoelasticity, 106
Maximum packing fraction, of suspension,
393
Maxwell element, 14,43,52
(see also Maxwell model)
Maxwell model, 52
exponential shear predicted by, 226
(see also Generalized Maxwell model)
Maxwell model parameters
determination of, 70
strain-dependent, 188
MECA Creep Rheometer, 286
Mechanical loss angle, 60
for several polymeric liquids, 61

657

Melt drawing (see Extrudate' drawing)


Melt fracture
in oscillatory shear, 222
(see also Extrudate distortion; Wall slip)
Melt I (LDPE), 133-135, 138,.160, 188, 190,
198, 203, 213, 214, 216,249
Melt index, 159; 327
effect of evolved gas; 577
measurement of, 328
on-line measurement of, 559
role of in blow molding, 528
role of in extrusion, 462
role of in injection molding, 501, 506, 507
use of for quality control,597
Melt index elasticity test; 215'
Melt spinning
use of birefringence in, 362
role of viscosity in, 586
use of for quality control, 598
(see also Extrudate drawing)
Melt strength, 247, 248
related to parison swell, 528
role of in extrusion casting; 554
role of in film blowing, 549
use of for quality. control; 598
Memory, of viscoelastic material, 22
Melting point, 574
related to Tq , 620
Memory function
from Lodge's network theory, 117
nonlinear, 131, 190, 201
relationship with relaxation time, 115
relationship with spectrum function, 116
separability of, 130, 132
strain dependent, 128, 129
Mesogens, 424
Mesophase, 424
Metravib, 245, 626
Metrilec Sari, 329, 626
Microdielectrometer, 417
Mitech Corporation, 623
Mixing, in extruders, 480-484
Modified Rouse theory (see Rouse model)
Modulus of rigidity, 11
(see also Shear modulus)
Moisture, absorption of by polymers, 579
Moldability tests, 502
Molecular dynamics, 107, 108
Molecular models for nonlinear
viscoelasticity, 146
Molecular motions, 367
Molecular orientation, 214
in blown film, 533
Molecular stretching in simple extension, 238
Molecular structure, 366

658

SUBJECT INDEX

Molecular theories
for entangled melts, 79
overview of, 365
prediction of linear behavior by, 74
Molecular weight
effect of on viscosity, 368
average, between entanglements, 51
Molecular weight distribution (MWD)
effect of on steady state compliance, 373
effect of on viscosity curve, 376-381
description of, 607
determination of from rheological
measurements, 574, 585
effect of on extensional flow properties,
259
effect of on tensile viscosity, 254
log normal, 373, 378, 610
measurement of, 611
most probable, 610
Monohole flow test, 414
Monsanto Instruments, 626
Morphology, of a blend, 402
Most probable MWD, 610
MTS, 294, 626
Mueller matrix, 352, 354
Multiphase systems, 390
Multistep strain, 191, 193
MWD (see Molecular weight distribution)
Nametre, 626
Navier-Stokes Equation, 40
Neck-in, 554
Net stretching stress, 31
for biaxial extension, 260
for step strain of linear material, 46
(see also net tensile stress)
Net tensile stress, 43, 58, 238
Newton's law of action and reaction, 27
Newton's second law of motion, 37, 39
Newtonian fluid, 2, 3, 36, 41
oscillatory shear of, 62
as special case for power law, 161
dilute suspension in, 390
entrance pressure drop of, 321
in simple shear, 36
normal stress differences, 30
pressure drop in orifice, 322
simple extension of, 37
slit flow of, 308
viscosity of, 11
viscosity of from tube flow data, 301
Nip rolls, 531, 532, 541
No-slip assumption, 158
Nonlinear viscoelastic behavior, 104-107, 179
classification of, 145
rheometers for study of, 270

Normal stress, 6
Normal stress differences, 31, 154, 155
for Newtonion fluid in simple shear, 37
for Newtonion fluids, 30
for rubberlike liquid, 118
second, prediction of by Doi-Edwards
theory, 190
Normal stress relaxation functions
for rubberlike liquid, 119
Nuclear magnetic resonance (NMR), 619
Number average molecular weight, 609
Nylon
MWD of, 610
On-line rheometers
applications of, 557, 563, 564
capillary type, 558
compared to in-line rheometers, 558
types of, 558
Operating diagrams
for blow molding, 592
for extruders, 455-459
Optical axes, 349
Orientation, of molecule, 20, 227
Orifice flow, pressure drop for, 322
Oscillatory shear
effect of fluid inertia on, 276
large amplitude, 219, 222, 223
limits of linear behavior, 222
of rubberlike liquid, 121
small ampli tude, 60
superposed on steady shear, 218
use of to probe relaxation, 219
p-azoxyanisole, 429
p-hydroxybenzoic acid, 425
Parallel disk flow, as viscometric flow, 157
Parallel disk rheometer, use of to study
curing reaction, 417
Parison formation, 590
of engineering resins, 522
modelling of, 521
role of rheological properties in, 525
Parison inflation
as extensional flow, 237
modelling of, 522
role of rheological properties in, 527
Parison programming, 510, 512
Parison sag, 512
as example of uniaxial extension, 236
combined with swell, 520
of engineering resins, 522
measurement of, 520
related to operating conditions, 526

SUBJECT INDEX

related to resin properties, 526


related to rheological properties, 519
Parison swell, 512, 513
combined with sag, 520
cuff effects, 514
definitions of, 513
effect of die design on, 515
effect of flow rate on, 515
effect of MWD on, 515
effect of temperature on, 514
effect of die design on, 516, 517
importance of, 514
isotropic, 515
isotropic versus anisotropic, 516
measurement of, 517, 526
prediction of, 519
related to capillary swell, 518, 527
related to stress ratio, 519
relationship between different types, 513,
516
time dependence of, 514
PDMS (see Polydimethylsiloxane)
PET (see Polyethylene terephthalate)
Pillow mold (see Parison swell, measurement
of)
Pinch-off mold (see Parison swell,
measurement of)
Pipkin diagram for oscillatory shear, 220, 222
Planar extension, 234
as a strong flow, 265
comparison with exponential shear, 151
material functions of, 264
methods for studying, 264
of LDPE, 265
rate of deformation tensor for, 263
role of in film blowing, 543
velocity distribution for, 263
Plasticity, 18
Plastication, in extruders, 442
Plasticizers, 619
Plateau compliance, 58
effect of polydispersity on, 60
Plateau modulus, 50, 51
in Doi-Edwards theory, 83
related to steady state compliance, 372
Plateau zone 50, 66
Pleating, in blow molding, 521, 590
PMMA (see polymethylmethacrylate)
Poiseuille Flow, as viscometric flow, 156
(see also Capillary flow)
Polarizability, 358
Polarization diagrams, 348
Poly(n-octyl methacrylate), storage
compliance of, 90
Poly-p-benzamide, 425 ,
Poly-p-phenylene teraphthalamide, 425

659

Polybutadiene
normal stress differences for, 191
stress growth coefficients for, 199
Polycarbonate, 574
Polydimethylsiloxane (PDMS)
cross-linking of, 420
damping function for, 140
entrance flow of, 319, 340
extrudate distortion of, 340
Polydispersity
effect of on relaxation spectrum, 98, 99
effect of on steady state compliance, 86
(see also Molecular weight distribution)
Polyethylene, 95
molecular structure of, 95
radius of gyration of, 21
time-temperature superposition applied
to, 92
volume of the sphere occupied by, 21
(see also HDPE; LDPE; LLDPE)
Polyethylene terephthalate (PET), 384, 575
absorption of water by, 579
blow molding of, 521, 523
crystallization of, 619
glass transition of, 619
injection blow molding of, 523
MWD of, 610
Polyisobutylene, 219
bidirectional start-up flow of, 199
cessation of steady simple shear, 201
damping function for, 136
extensional flow of, 266
Polyisobutylene solution
damping function for, 140
Polymer laboratories, 627
Polymer molecules, why elastic, 19
Polymer solutions, 19
Polymerization reactions, 576
Polymethyl methacrylate (PMMA), 162, 370
extensional flow of, 259
Polypropylene
extrudate distortion of, 339
extrudate swell of, 334
film blowing of, 547
MWD of, 561, 611
parison swell of, 514
use of rheometer for, 561
Polystyrene, 162, 190, 370, 598
compliance of, 93
damping function for, 135, 137, 140
extensional flow of, 259
extrudate distortion of, 339
nonlinear creep compliance of, 208
nonlinear relaxation modulus of, 188
normal stress differences for, 191
relaxation modulus of, 183

660

SUBJECT INDEX

Polystyrene (cont' d.)


relaxation spectrum of, 98
shear stress growth coefficient of, 196
start-up of steady simple shear, 195
storage and loss moduli of, 95-97
tensile creep, 243
time-temperature superposition of, 93
type II nonlinear behavior of, 196
viscosity curves for, 163
Polystyrene solution
damping function for, 140
Polyurethane
cross-linking of, 420
Polyvinyl chloride (PVC), 434, 584
die flow of, 510
Position vector, 32
Potential function
of factorable BKZ model, 128
Power law for viscosity, 160-162, 166, 376
for liquid crystal polymers, 435
in capillary flow, 301
in injection molding, 493
use of to model flow in mandrel die, 538
use of to model die flow, 510
use of to model extrusion, 464
Pressure, 29, 30, 39
Pressure, effect of on viscosity, 385
Principal strain axes, 233
Principal stresses, extensional flow, 238
Principal stretching stress (see Net tensile
stress)
Process control
manipulated variables for, 564
use of rheometers for, 564
Processability 180, 181, 584
Processing aids, use of in film blowing, 539
Pseudoplasticity (see Shear thinning
behavior)
PVC (see polyvinyl chloride)
Quality control, 104, 159, 594-599
use of solution viscosity for, 615
selection of rheometer for, 596
for blow molding resins, 528
Rabinowitch correction, for capillary flow,
303, 325, 326
Radius of gyration, of branched molecule,
387
Random walk model, 357
Rate of deformation, for simple shear, 35
Rate of deformation tensor, 34
for extensional flows, 233
Reaction injection molding (RIM), 411
rheometers for study of, 416

Reactive extrusion, use of on-line rheometers


for, 565
Recoil
during creep experiment, 57
during start-up flow, 214
methods of measuring, 286
(see also Ultimate recoil; Tensile recoil)
Recoil function, 57
for typical melts, 58-59
nonlinear, 211
Recoverable compliance (see Recoil
function)
Recoverable shear (see Ultimate recoil)
Recoverable shear strain (see Recoil)
Recoverable strain, 257
at high strains, 253
of polystyrene, 259
(see also Ultimate recoil; Ultimate tensile
recoil; Tensile recoil)
Reduced viscosity, 613
Reduction in shear rate test, 205
Reentanglement time
for interrupted shear, 204
for reduction in shear rate, 205
Reference configurations, 112
Reference state, 7, 112
Refractive index
as tensor quantity, 349, 357
definition of, 346
imaginary, 349
related to stress, 358
Relative viscosity, 613
of a suspension, 393
Relaxation modulus, 44, 48, 51
Doi-Edwards prediction of, 83
for Maxwell element, 52
empirical equations for, 51
effect of polydispersity on, 51
nonlinear, 132, 133, 183, 189
of cross-linking material, 420
of typical polymeric materials, 49
power-law form of, 54
relationship to memory function, 116
relationship with storage and loss moduli,
64
separability of, 132, 184
type I nonlinear behavior, 186
type II nonlinear behavior, 186, 188, 193
Relaxation spectrum, 54
and Maxwell parameters, 72
from experimental data, 55
of polystyrene, 98
relationship to storage and loss moduli, 64
shear rate-dependent, 167
strain-dependent, 188
truncation of at high shear rates, 167, 168

SUBJECT INDEX

Relaxation strengths
and recoil, 216
from tensile viscosity, 257
shear-rate dependent, 202
Relaxation time
and tensile stress growth function, 125
longest, 169
longest, from Doi-Edwards theory, 84
longest, Rouse, 77, 147
of a Maxwell element, 16
role of in oscillatory shear, 220
effect of temperature on, 87
from modified Rouse theory, 76
of generalized Newtonion model, 71
Reptation, 81, 83, 85, 86, 146
Resin optimization, 588-595
Resin selection, 585
Rest time, for interrupted shear, 203
Retarded elasticity, 14
Retraction time, 148, 184
Reynolds number, 41
Rheo-optics, 345
Rheological testing, 568-577
Rheology
aspects of, 2
definition of, 1
Rheometer compliance (see Compliance, of
rheometers)
Rheometers
automation of, 572
concentric cylinder, 285
cone-plate, 277
advantages of, 278
basic equations for, 278
edge effects in, 280-281
sources of error for, 279
controlled stress, 286
drag flow, sources of, 270
eccentric rotating disks, 284
extensional, 242, 244, 245
industrial use of, 567
manufacturers of, 622
parallel disk, 283, 286
sandwich, 291
selection of, 567-573
sliding cylinder, 294
sliding film, 293
sliding plate, 287
sliding plate: basic equations for, 288
design of, 291
end and edge effects, 289
uses of, 290
sliding surface, 269
use of to measure N\, 282
(see also Torque rheometers; Capillary
rheometers; Slit rheometers)

661

Rheometries, 241, 245, 286, 560, 561, 564,


627
Rheopexy, 17
Rheoplast, 329
Rheoprocessor, 561
Rheovibron, 416
RIM (see Reaction injection molding)
Rise time, in step strain, 181
Rod climbing (see Weissenberg effect)
Rosand Precision Limited, 627
Rotary clamp, 243
use of to measure melt strength, 247
use of with HDPE, 258
use of to generate biaxial extension, 262
Rouse model, 146
for dilute solutions, 74
and polydispersity, 78
Bueche modification of, 75
nonlinear effects, 146
prediction of linear viscoelastic properties,
77
Rouse theory, 371, 377
Rubber, 2, 3, 7, 13, 25
optical properties of, 350
polarization of, 350
refractive index of, 350
Rubberlike liquid, 201
and spike strain test, 193
cessation of steady shear, 121
comments on, 126
deficiencies of, 126
definition of, 115
exponential shear of, 228
for any simple shear deformation, 115
generalization of, 130
in oscillatory shear, 121
in simple extension, 123
in simple shear flows, 118
in steady simple shear, 119
in step shear strain, 119
normal stress differences for, 118
recoil of, 122
ultimate recoil following steady simple
shear, 216
Runners
design of, 493
flow balancing in, 493
flow in, 492
heated, 492
(see also Injection molding)

Sag (see Parison sag)


Sandwich rheometers (see Rheometers)
Scalar invariants of the Finger tensor, 129

662

SUBJECT INDEX

Scalar invariants of the Finger tensor (cant' d.)


dependence of damping function on, 130,
140

for simple extension, 131


for simple shear, 131
for simple extension, 114
Scalar invariants of a vector, 113
Schiimmerapproximation for wall shear rate
in capillary flow, 304, 326
Screen pack, 531
Screening of resins, 573, 593
Second normal stress growth coefficient, 195
Second normal stress difference, 154,
and BKZ equation, 129
prediction of by Doi-Edwards equation,
149

Normal stress differences)


Second Normal Stress Coefficient
definition of, 155
Second normal stress decay function and
coefficient, 200
Secondary flow, in cone-plate rheometers,
(see also

279

Seiscor, 313, 559, 560, 627


Seiscor/Han Rheometer, 313, 560
Separability
of memory function, 128, 145
of relaxation modulus, 184
Sharkskin, 337-338
(see also Extrudate distortion)
Shear compliance, 11
Shear creep compliance (see Creep
compliance)
Shear-free flows, 233
(see also Extensional flows)
Shear modification, 205
effect of on extensional flow, 540
in blow molding process, 511
in film blowing, 540
role of branching in, 388, 540
Shear modulus, 11
(see also Relaxation modulus)
Shear rate, 10
in simple shear, 153
Shear refining (see shear modification)
Shear relaxation modulus (see Relaxation
modulus)
Shear strain, 9, 10
in simple shear, 153
Shear stress, 6, 28
Shear stress decay coefficient, 73, 200, 201
Shear stress decay function, 199
Shear stress growth coefficient, 72, 73, 195
as a function of shear strain, 112
related to tensile stress growth coefficient,
231,239

Shear stress growth function, 194


(see also Shear stress growth coefficient)
Shear stress transducer
use of in an extruder, 562
use of in an in-line rheometer, 562
Shear thickening behavior, of liquid crystal
polymers, 437
Shear thinning behavior, 12
effect of MWD on, 376
Shear Wave Propagation, in drag flow
rheometers, 275
Sheet extrusion, 247, 552
edge bead, 553
draw resonance, 554
neck-in, 554
role of melt strength, 554
Sheet inflation
use of to generate planar extension, 264
use of to study biaxial extension, 261
Shift factor, for time-temperature
superposition, 88, 89, 94, 173
as a function of temperature, 93
for viscosity, 169
Shimadzu Scientific Instruments, 628
Simple extension, 7, 29, 31
definition of, 234
displacement functions for, 606
displacement gradient tensor for, 606
infinitesimal strain tensor for, 35
of Newtonion fluid, 37
shear rate tensor for, 35
velocity distribution, 237
Simple shear, 6, 9, 28
coordinate system for, 233
displacement function for, 601-602
invariants of Finger tensor for, 113
rate of deformation tensor for, 36
rheometers, 269
stress components for, 29, 154
(see also Steady simple shear)
Simulation of melt processes, 587
Sinusoidal shear (see Oscillatory shear)
Sliding cylinder flow, as viscometric flow, 157
(see also Rheometers)
Sliding plate rheometers (see Rheometers)
Slip flow (see Wall slip)
Slit flow
advantages of, 307
as viscometric flow, 156
determination of wall shear rate in, 309
of Newtonian fluid, 308
shear stress distribution in, 307
use of to determine Nt, 309
wall shear stress in, 308
(see also Slit rheometers; Entrance effects;
Hole pressure; Exit pressure)

SUBJECT INDEX

Slit rheometers
for on-line use, 559
(see also Slit flow; Hole pressure; Exit
pressure)
Slot casting (see Sheet extrusion)
Solid body motion, 604
Solid body rotation, 604n
Solution viscosity, 613
Specific viscosity, 613
Spike strain test, 193
Spiral mandrel die, design of, 538
Spiral mold test
for cross-linking systems, 414
for injection molding resins, 503, 505
Squeezing flow
use of to evaluate epoxy resin, 414
(see also Lubricated squeezing flow)
Stability of polymer
importance of, 577
measurement of, 578
Stagnation flow, 545, 546
Start-up of planar extensional flow, 264
Start-up of steady simple extension, 239
linear response, 74
of a rubberlike liquid, 124
Start-up of steady simple shear
linear response in, 72
nonlinear behavior, 194
Steady simple shear
definition of, 155
importance of, 153
linear viscoelastic behavior, 47
velocity distribution for, 36
Steady state compliance, 55, 58, 65, 70, 207
effect of MWD on, 57, 60, 86, 372, 380
effect of temperature on, 374
effect of branching on, 388
of blends, 372
related to molecular weight, 371
related to N!, 121, 374
related to plateau modulus, 372
related to viscosity, 377
from Doi-Edwards theory, 84
for linear behavior, 56
from modified Rouse model, 77
ofLDPE,254
Steady state tensile compliance, 241
Step shear in extension, 123
Step shear strain, 181, 183
use of to probe relaxation, 219
(see also Relaxation modulus)
Storage and loss compliances, 69
for typical linear polymer, 70
Storage and loss moduli
use of to determine Maxwell parameters,

100

663

calculation of, 220


calculation of by data acquisition system,
572
definition of, 62
dependence on strain or rate amplitude,
222
effect of molecular weight on, 95
for linear polymers, 67
from modified Rouse model, 77
low-frequency limiting behavior of, 65
of cross-linking material, 420
of typical molten polymers, 66
relationship to molecular weight
distribution, 68, 96
(see also Complex modulus)
Stored elastic energy, 214
Strain, 6, 7, 605
finite measures of, 108
nonlinear measures of, 131
(see also Strain tensor)
Strain hardening (see Extension thickening)
Strain rate (see Rate of deformation)
Strain softening behavior, definition of, 126
(see also Extension thinning)
Strain tensor
Cauchy strain tensor, 111
Finger strain tensor, 111
for simple extension, 35
for simple shear, 34
infinitesimal, 31, 33, 34, 105, 111
of Seth, 130
relative, 112
Strength of the network at the gel point, 420
Stress, 3, 5
(see also Stress tensor)
Stress growth coefficient (see shear stress
growth coefficient; Tensile stress growth
coefficient)
Stress growth function (see shear stress
growth coefficient; tensile stress growth
coefficient)
Stress-optical coefficient, 358
Stress overshoot, in steady simple shear, 196
Stress ratio
and the recoverable shear, 122
and ultimate recoil, 214, 216
related to parison swell, 519
Stress relaxation, 16, 43
in molten polymers, 50
of elastomer, 48
(see also Relaxation modulus)
Stress tensor, 25
components, 27
components of simple shear, 27, 154
for extensional flows, 238
for simple extension, 29

664

SUBJECT INDEX

Stress tensor (cont'd)


in fluid at rest, 30
meaning of indices, 27
relationship to surface stress vector, 26
sign convention for, 27
symmetry of, 28, 154
Stress-optical relation, 357-359
failure of, 363
Stretching flow, 232
Strong flow, 20, 150,265
exponential shear as, 225
simple extension as, 238
birefringence in, 360
Structural Time Dependency, 16
Structure
of a melt, 217
of a fluid, 16
Substantial derivative, 39
Superposed deformations, 217, 219
Superposed steady and oscillatory shear, 218
Surface force, 4, 5, 25
Surface stress vector, 25
Surface tension, effect of on tensile
measurement, 244
Suspensions, 17
(see also Filled melts; Viscosity)
Swell (see Extrudate swell)
Temperature distribution, in drag flow
rheometers, 274
Tensile creep compliance
nonlinear, 240
of polystyrene, 259
Tensile flow (see Simple extension)
Tensile recoil, 241
ofLDPE,249
prediction of, 257
Tensile relaxation modulus, 43, 238
Doi-Edwards prediction of, 148
Tensile start-up flow, 254
Te~ile stress,S, 24
(see also Net tensile stress)
Tensile stress decay coefficient, 240
linear behavior, 73
Tensile stress growth function
for rubberlike liquid, 124, 125
of LDPE, 251
Tensile stress growth coefficient
at high strains, 253
definition of, 239
linear behavior, 73
ofHDPE,258
of LDPE, 250, 251
of polystyrene, 259
related to shear stress growth coefficient,
231
Tensile stress relaxation, 133

Tensile viscosity
effect of molecular weight on, 255
effect of MWD on, 254
from entrance pressure drop, 323-324
ofHDPE,258
of LDPE, 251, 255
of rubberlike liquid, 124
use of for quality control, 598
(see also Apparent extensional viscosity)
Tensor, 22-23
as operator, 24
components of, 24
notation for, 25
Terminal relaxation time, 53
related to steady state compliance, 57
Terminal zone, of viscoelastic behavior, 50,
58,66
Testing Machines Inc., 628
Thermal stability, 577
Thermogravimetric analysis, 577
Thermorheologically simple behavior, 87
Thermosetting materials, 411
rheological testing of, 414
Thickness swell (see Parison swell)
Thixotropic loops, 221
Thixotropy, 17
Time constant in viscosity equation, 161-163
Time-temperature superposition, 86, 87, 94,
96
and viscosity function, 169
creep recovery, 213
failure of, 92
nonlinear viscoelasticity, 188
of nonlinear creep time, 209
start-up of steady shear, 215
ultimate recoil, 215
Time-temperature-transformation, 411
Time-Temperature Instruments, 286, 628
Tinius Olsen, 628
Torque rheometers, 287
use of to study stability, 578
Torsional braid analysis (TBA), 416
Torsional impregnated cloth analysis (TICA),
416
Toyo Seiki, 628
Transducers
compliance of, 271
for shear stress, 292, 293
Transient shear tests, usefulness of, 228
Transition zone, of viscoelastic behavior, 50,
58
Triple-step shear strain test, 194
Tube flow (see Capillary flow; Channel
flow)
Tube model (see Doi-Edwards theory)
Tube renewal, 85
Twin-screw extruders, 485-489

SUBJECT INDEX

Type I nonlinear behavior, 186


Type II nonlinear behavior, 186, 188, 193
Ultimate recoil, 57
during start-up flow, 73
following steady shear, 214
following steady simple shear, 215
for a crosslinked material, 123
nonlinear, 211
of LDPE, 213, 251
of rubberlike liquid, 122
relationship to loss angle, 216
relationship to stress ratio, 216
time-temperature superposition, 213, 215
(see also Recoil; Tensile recoil; Ultimate
tensile recoil)
Ultimate recoil function, 252
nonlinear, 211
Ultimate tensile recoil, 241
measurement of, 244
of LDPE, 251
Uniaxial extension (see Simple extension)
Unit tensor, 111
Vector, 23
Vectra,425
Venturi flow, 546, 547
Viscoelasticity, 42
of a rubber, 13
Viscometric flow
and low frequency oscillatory shear, 220
examples of, 156
definition of, 155
Viscometric functions, 155, 195
Viscosity, 11
definition of, 12, 155
dependence of on temperature, 169, 381,
582-585
dependence of on molecular weight
distribution, 374
dependence of on pressure, 384
dependence on shear rate and molecular
weight, 166
dependence on shear rate, 159, 374-381
effect of filler concentration on, 390
effect of filler particle asymmetry on, 394
effect of filler particle size distribution on,
394
effect of filler volume fraction on, 392
effect of molecular weight on, 164-167
factors it depends on, 159
of a Newtonion fluid, 11
of a suspension, 391
of a suspension, effect of shear rate on,
395
of a concentrated suspension, 392
of fiber reinforced melts, 393

665

of immiscible blends, 401


of molten polymers, 158
of pre-gel liquid, 418
pressure-independent plot of, 169
relationship to damping function, 168
shear rate for onset of shear thinning, 377
(see also Zero shear viscosity)
Viscosity average molecular weight, 611, 615
Viscous dissipation (see Dissipation)
Viscous heating
in capillary rheometers, 326
in rheometers, 274
(see also Dissipation)
Viscous stress (Extra stress), 31
Vitrification, during cross-linking, 411
Voigt body, 13, 48
Volume fraction, effect on viscosity, 391
Wagner's equation, 130, 193-194
and exponential shear, 227
generalization of, 266
not a constitutive equation, 146
Wall slip, 158
in capillary flow, 305
related to extrudate distortion, 340
Weak flow (see Strong flow)
Weber Number, for immiscible blends, 403
Weight average molecular weight, 608
Weight swell (see Parison swell)
Weissenberg effect, 105, 275, 285
Weissenberg number
for channel flow, 331
for oscillatory shear, 221
Weld line
healing of, 495
in injection molding, 495
reducing the effects of, 496
role of temperature, 496
in blow molding, 510
WLF equation, 89, 382-383, 584
Yield stress, 18, 397
measurement of, 397
methods of measuring, 286
of a suspension, 395, 397, 398
Yo-yo model, 360
Z-average molecular weight, 609
Zero-shear viscosity, 12
dependence of on temperature, 89
from modified Rouse theory, 76
relationship to relaxation modulus, 73
Doi-Edwards theory, 84
effect of molecular weight on, 80, 164-167
of blend, 166
related to relaxation modulus, 47
related to molecular weight distribution,
368-370

You might also like