Sei sulla pagina 1di 491

NONLINEAR PHENOMENA IN FLOWS

OF VISCOELASTIC POLYMER FLUIDS

NONLINEAR PHENOMENA IN FLOWS


OF VISCOELASTIC POLYMER FLUIDS

A.I. LEONOV
Institute of Polymer Engineering,
College of Polymer Science and Polymer Engineering,
The University of Akron, Akron, Ohio, USA

AND
A.N. PRO KUNIN
Institute of Problems in Mechanics,
Academy of Science of Russia, Moscow, Russia

SPRINGER-SCIENCE+BUSINESS MEDIA, B.V.

First edition 1994

1994 Springer Science+Business Media Dordrecht


Originally published by Chapman & Hall in 1994
Softcover reprint of the hardcover 1st edition 1994
Typeset in 10 on 12pt Times by Interprint Limited, Malta

ISBN 978-94-010-4548-3
ISBN 978-94-011-1258-1 (eBook)
DOl 10.1007/978-94-011-1258-1
Apart from any fair dealing for the purposes of research or private study, or criticism or
review, as permitted under the UK Copyright Designs and Patents Act, 1988, this
publication may not be reproduced, stored, or transmitted, in any form or by any means,
without the prior permission in writing of the publishers, or in the case of reprographic
reproduction only in accordance with the terms of the licences issued by the Copyright
Licensing Agency in the UK, or in accordance with the terms of licences issued by the
appropriate Reproduction Rights Organization outside the UK. Enquiries concerning
reproduction outside the terms stated here should be sent to the publishers at the London
address printed on this page.
The publisher makes no representation, express or implied, with regard to the accuracy
of the information contained in this book and cannot accept any legal responsibility or
liability for any errors or omissions that may be made.
A catalogue record for this book is available from the British Library

Contents

Preface
Acknowledgement
Chapter 1 Constitutive Equations with a Recoverable Strain
Tensor as an Internal Parameter
1.1
Introduction
....
. . . .
1.2
Decomposition of the total strain into recoverable and
irreversible components and their relation to the
strain rate
...........
1.3
Dependences of stress on elastic (recoverable) strain
and temperature
. . . . . . . .
1.4
Balance of mechanical energy and dissipation
1.5
General dependences of the irreversible strain rate and the
dissipative function on the elastic strain and temperature
1.6
Specifications of the elastic potential and the irreversible
strain rate
........
....
1.7
On the thermodynamic stability and positive definiteness
of the dissipative function for the model specified with
. . . . . . . .
eqns (1.51) and (1.55)
1.8
Nonlinear multimodal Maxwell constitutive equations
1.9
Multimodal approach in the limit case of linear
viscoelasticity
. . . . . . . .
1.10 On the rheological parameters in nonlinear multimodal
Maxwell constitutive equations
....
1.11 On the simplest version of the viscoelastic model with
relaxation and retardation
. . . .

Xlll

xvii

5
10
12
14
16
21
23
25
28
30

vi

Contents

Chapter 2 Other Constitutive Equations for Elastic Liquids


2.1
2.2

2.3

Chapter 3

3.1
3.2
3.3

3.4

3.5

Introduction
Some popular nonlinear viscoelastic constitutive
equations derived from continuum mechanics
2.2.1 Constitutive equations of differential type
2.2.2 Constitutive equations of integral type
2.2.3 Rational mechanical approach
Some physical models of polymeric liquids
2.3.1 Statistical mechanics of rubber elasticity
2.3.2 Statistical mechanics of dilute polymer solutions
2.3.3 Microscopic models of polymer melts and
concentrated polymer solutions

Analyses of Simple Constitutive Equations for


Viscoelastic Liquids
Introduction
Thermodynamic derivation of general Maxwell-like
constitutive equations
Some instabilities in viscoelastic Maxwell-like and single
integral constitutive equations
3.3.1 Hadamard instability: method of 'frozen
coefficients'
3.3.2 Equivalence between Hadamard and
thermodynamic stabilities for Maxwell-like
constitutive equations with convected derivatives
3.3.3 Constraints imposed on possible forms of elastic
potential by Hadamard (thermodynamic) stability
3.3.4 Hadamard stability of single integral constitutive
equations
3.3.5 Positive definiteness of the configuration tensor:
'Dissipative' instability
Examples
3.4.1 Upper-convected Maxwell model
3.4.2 Dumbbell-spring model with finite extensibility
3.4.3 Phan-Thien/Tanner upper-convected models
3.4.4 Larson model
3.4.5 Giesekus model
3.4.6 Oldroyd-Lodge model
3.4.7 Wagner model I
3.4.8 Wagner model II
3.4.9 Kaye-BKZ model with Mooney-Rivlin potential
(eqn (1.46))
Discussion
3.5.1 The importance of tre p for the formulation of
Maxwell-like constitutive equations
3.5.2 Hadamard and dissipative instabilities

32
32

33
33
37
39
43
43
45
49
54
54

55
59
60
63
65
67
68
70
70
71
71
72
72
73
73
73
73
74
74
75

Contents

Chapter 4 Experimental Methods in the Rheology of Viscoelastic


Liquids
4.1
Introduction
4.2
Simple shear tests
4.2.1 General relations
4.2.2 Parallel plate device
4.2.3 Cone-plate rheometer: basic relations
4.2.4 Cone-plate: schemes of instruments and methods of
measurement
4.2.5 Methodical remarks
4.3
Disk-disk rheometer
4.4
Capillary rheometry
4.4.1 Basic elements and operational procedure
4.4.2 Determining flow curves in isothermal case
4.4.3 Entry and exit corrections in capillary flow
4.4.4 Dissipative heating
. . . . . .
4.4.5 Evaluation of wall slip
.......
4.5
Experimental methods in the uniform extension of liquid
polymers
. . . .
4.5.1 Basic relations
4.5.2 Schemes of instruments
4.5.3 Methodological aspects
4.5.4 Defects in schemes of devices and errors in
measurements
........
4.6
Experimental methods in non-uniform extension of liquid
polymers . . . . . . . . . . . . . . . . .
4.6.1 Weakly non-uniform extension
......
4.6.2 Schemes of devices for the drawing of elastic liquids
with a free surface
4.6.3 Drawing out of a capillary
4.6.4 Some methodological observations
4.7
Methods of flow birefringence
Chapter 5

5.1
5.2
5.3
5.4
5.5

Theoretical and Experimental Investigation of Shear


Deformations in Elastic Polymeric Liquids
Introduction
. . . . . . . . .
Constitutive equations for simple shear
Shearing in the linear region of deformations
5.3.1 Experimental results
5.3.2 Theoretical considerations
Evaluation of rheological parameters
Homogeneous shearing under a given shear rate (start-up
flow and stress relaxation)
5.5.1 Experimental results
5.5.2 Theoretical considerations
5.5.3 Discussion of results

vii

78

78
79
79
80
81

84
87
92
94
94
94
95
96
96
97
97
100
102
103
105
105
106
108
109
109
113
113
114
117
117
120
122

124
124
129
132

viii

Contents

5.6
5.7

5.8

5.9
5.10
5.11

Parallel superposition of oscillations on steady shear flow


5.6.1 Experimental results
5.6.2 Theoretical considerations
Orthogonal superposition of oscillations on steady shear
flow
5.7.1 Experimental results
5.7.2 Theoretical considerations
Shear deformations under given shear stress: start-up flow
and elastic recovery
5.8.1 Experimental results
5.8.2 Theoretical considerations
Isothermal flow between two rotating disks
5.9.1 Experimental results
5.9.2 Theoretical considerations
On calculations of shear deformations using generalized
power-like elastic potential
Qualitative features of simple shearing as predicted by
some other Maxwell-like constitutive equations
5.11.1 Giesekus model
5.11.2 Larson model
5.11.3 Discussion of blow-up shear instabilities in
Giesekus and Larson constitutive equations
5.11.4 A simple viscoelastic Maxwell-like constitutive
equation

Chapter 6 Experimental and Theoretical Studies of Uniaxial


Uniform Extension of Polymeric Liquids
6.1
6.2
6.3

6.4
6.5

6.6
6.7

Introduction
Studies of regimes of linear extensional deformations
Experimental studies of extensional deformation in a
nonlinear region
6.3.1 Evolution of elongation strain
6.3.2 Dependences of stress and strain rates on elastic
(recovery) strain
6.3.3 Stress and strain relaxation
Constitutive equations for uniform extension
Theoretical treatment
6.5.1 Extension under a given constant strain rate
6.5.2 Extension under a given constant force
6.5.3 Extension under a given constant stress
6.5.4 Elastic recovery after uniaxial extension
Comparison of theoretical calculations with data
Qualitative features of simple extension as predicted by
some other Maxwell-like constitutive equations
6.7.1 Giesekus model
6.7.2 Larson model
6.7.3 Steady simple elongation for a new constitutive
equation

134
134
137
144
144
144
152
152
154
157
157
159
161
162
162
165
166
167

170
170
171
176
176
181
185
188
189
189
190
192
193
195
196
197
197
198

Contents

Chapter 7

7.1
7.2
7.3
7.4

7.5

On Hardening Phenomena in Flows of Polymeric Liquids


Introduction
Additional data on effective extensional viscosity
Hardening effects in extensional flows of polymeric
liquids
Modelling the hardening phenomena in extensional flows
of polymeric liquids
7.4.1 General remarks
7.4.2 Modelling of hardening as the relaxation transition
7.4.3 Modelling of hardening phenomena with a long
time and low elastic modulus relaxation mode
Modelling of hardening phenomena in simple shearing of
elastic liquids
7.5.1 Hardening as a relaxation transition
7.5.2 Modelling of hardening phenomena with a long
time and low elastic modulus relaxation mode

Chapter 8 Flows of Polymeric Viscoelastic Liquids in Channels


and Pipes
Introduction
8.1
8.2
General results for rectilinear steady flows of viscoelastic
liquids in long tubes with arbitrary cross-sections
8.3
Predictions of Maxwell-like constitutive equations for
steady flows in long tubes of arbitrary cross-sections
8.3.1 (I2 =0: existence of pure rectilinear flow
8.3.2 I(I 21 ~ min(1 (I 121, (I d: weak secondary flow imposed
on a strong rectilinear flow
8.3.3 Strong secondary flows
8.4
Poiseuille flows of viscoelastic liquids in capillaries and die
slits
8.4.1 General formulation
8.4.2 Steady Poiseuille flows
8.4.3 Hardening phenomena in steady capillary flows
8.4.4 Unsteady channel and capillary flows of viscoelastic
liquids
8.4.5 Enhancement of the flow rate using a pressure drop
with a pulsatile component
8.5
Entrance and exit steady flows
8.5.1 Experimental results
8.5.2 Numerical simulations
8.5.3 Comparison of numerical simulations with data
Chapter 9 Non-isothermal Flows of Polymeric Liquids
Introduction
9.1
9.2
Equation for temperature variations
9.3
Time-temperature superposition
9.4
Steady simple shear with dissipative heating
9.5
Hydrodynamic thermal explosion

IX

199

199
200
201
210
210
211
214
218
218
221

224
224
224
227
228
229
230
232
232
233
234
236
242
246
246
250
257
260
260
261
265
268
272

Contents

9.6

9.7
9.8

9.9

Steady shearing flow of polymeric liquids between rotating


disks with an account of dissipative heating
9.6.1 The scheme of apparatus
.... .
9.6.2 Theoretical analysis
.... .
9.6.3 Comparison of theoretical calculations with
experiments
. . . . .
Sealing of a rotating shaft with polymeric liquids
Non-isothermal capillary flows with an account of
dissipative heat generation
9.8.1 The problem formulation
9.8.2 Some results of a numerical study
9.8.3 An analytical approximation
9.8.4 Flows in long tubes
Non-isothermal elongational deformation of polymeric
liquids
.....
...... .
9.9.1 An example of quick thermal adaptation
9.9.2 Non-isothermal strain recovery

Chapter 10 Flows Close to Simple Shear and Simple Extension


10.1 Introduction
. . . .
10.2 Steady helical flow of viscoelastic liquids
10.2.1 General formulation
10.2.2 Expression for the flow rate
10.2.3 Results of theoretical calculations and a comparison
with experimental data
10.3 Viscoelastic flow in a disk extruder
10.3.1 Scheme of disk extrusion
10.3.2 Formulation of the problem
10.3.3 Solution of the problem
10.3.4 High intensity disk extrusion
10.4 On optimal geometry for generalized rotational flow of
viscoelastic liquids in thin gaps
. . . .
10.4.1 Basic relations and formulations
10.4.2 Isoperimetric problems of extremum of torque
10.4.3 On maximization of thrust
10.5 Weakly non-uniform extension of viscoelastic liquids
10.6 Certification of raw material for the polymer processing
industry
. . . . . . . . . . . .
10.7 On the stretching and swelling of an elastic liquid extruded
from a capillary die
..... .
10.7.1 Formulation of the problem
10.7.2 Steady capillary flow in region I
10.7.3 Stretching polymer extrudate outside the circular die
10.7.4 Experimental data and their comparison with
theoretical calculations
.....
10.8 Drawing polymeric liquids from a reservoir with a free
surface: open channel (elastic) siphon
10.8.1 Basic effect
....

274
275
276
278
280
282
282
285
287
290
290
291
294

297
297
298
298
300
302
305
305
306
307
308
310
310
313
315
319
325
326
327
328
329
331
338
338

Contents

xi

10.8.2 Theoretic model


10.8.3 Comparison with experiments
10.8.4 Generalizations
.....
10.9 Propagation of nonlinear waves along a viscoelastic bar
10.9.1 Fundamental equations
......
10.9.2 Weak discontinuities and shock waves in a
viscoelastic bar
10.9.3 Striking a viscoelastic bar of finite length against a
rigid obstacle
. . ..
10.10 The relationship between pure shear (planar extension) and
simple shear deformations for viscoelastic liquids

339
341
342
343
344

Chapter 11 Melt Flow Instabilities


11.1 Experimental data
11.1.1 Historical
11.1.2 Extrudate appearance
11.1.3 Superextrusion: second stable regime
11.1.4 Change of slope in flow curve
11.1.5 Hysteresis
. . . .
11.1.6 Oscillations
.....
11.1.7 Site for initiation of instability
11.1.8 Structure of polymer
11.2 Sharkskin and spurt instabilities
11.2.1 Sharkskin
.....
11.2.2 Spurt melt fracture
11.3 Mechanisms of melt flow instabilities
11.3.1 Viscoelastic criterion for the onset of unstable flow
11.3.2 Fluid mechanical instability
11.3.3 Fracture (discontinuity) of polymer melts in flow
11.3.4 Wall slip effects
. . . .
11.3.5 Other explanations and assumptions
11.3.6 A concept of polymer fluidity loss in intense
flows
. . . . ..
....
11.4 Sliding friction of crosslinked elastomers and wall slip of
polymer melts
.......
11.5 Self-oscillations of polymer melts in a rate-controlled
capillary rheometer
11.6 Other theoretical models

356

Additional Problems in the Rheology of Polymeric Fluids

396
396
400

Chapter 12

12.1
12.2

12.3
12.4

Strength of polymer melts under extension


... .
On detachment of polymers from walls
.... .
12.2.1 Detachment of polymer from capillary wall under
stretching of extrudate: raw resonance
12.2.2 Detachment of a deformed polymer from the wall
under the action of a low molecular fluid
Capillary break-up of polymeric jets
Flow-induced crystallization

346
348
353
356
356
357
359
360
364
365
368
369
370
370
372
379
379
'380
381
381
381
382
383
388
393

400
401
402
404

xii

Contents

12.5
12.6

Polymer degradation
. . . .
12.5.1 Thermal degradation and oxidation
12.5.2 Mechanical degradation and oxidation
Problems in the rheology of other polymeric systems
12.6.1 Rheology of filled polymers
12.6.2 Rheology of polymer blends
12.6.3 Rheology of liquid crystalline polymers
12.6.4 Chemorheology

405
406
407
407
408
408
409
409

Appendices
Al
Kinematics of continuum
ALl Eulerian and Lagrangian descriptions
Al.2 Basis vectors and tensors
Al.3 Strain gradient tensors
A1.4 Cayley polar decomposition
Al.5 Strain measures
Al.6 Invariants of tensors and Hamilton-Cayley identity.
D(ensity)
. . . . . . . . .
Al.7 Scalar functions of tensors and their tensor
derivatives
. . . . . . . .
Al.8 Strain rate and vorticity tensors
Al.9 Evolution equation for strains
ALlO Continuity equation
A brief introduction into non-equilibrium thermodynamics
A2
A2.l Conservation laws: the local formulation of the
First Law of Thermodynamics
....
A2.2 Local equilibrium assumption: Gibbs' relation. A
local formulation of the Second Law of
Thermodynamics
......... .
A2.3 Expressions for entropy production, entropy flux
......... .
and heat capacity
A2.4 Generalized thermodynamic forces and fluxes
Component-wise expressions for basic equations
A3
A3.l The equations of momentum balance and
continuity for incompressible media
A3.2 Equation for temperature variations
A3.3 Component-wise expressions for the velocity
gradient tensor and the upper-convected time
derivative of recoverable strain tensor

410
410
410
412
413
414
415

References

434

Index

464

416
418
420
421
422
423
423
423
425
425
426
426
428
429

Preface

This monograph presents theoretical and experimental studies of flows of


elastic liquids. Falling into this category are particularly the melts and
concentrated solutions of such flexible-chain polymers as polyethylene,
polyisobutylene and polypropylene, all of which are widely used in
polymer processing. These polydisperse polymers vary greatly, from
batch to batch, in their mechanical properties and 20% variation in a
property is believed to be good enough.
All recent books l - 7 devoted to the rheology of polymers do not answer
the question of which constitutive equations should be used for solving the
fluid mechanic problems of polymer processing in the usual case of an
appreciable nonlinear region of deformation where nonlinear effects of
shear and extensional elasticity are very important. Viscoelastic constitutive equations cited commonly (see, e.g. Refs 5 and 6) do not describe
simultaneously even the simplest cases of deformations, viz. simple shear
and uniaxial extension. Moreover, some of them are internally inconsistent and sometimes display highly unstable behaviour in simple flows
without any fundamental reasons. Even more respected molecular approaches are not free from these defects.
In the present book, the authors attempt to set forth the reading
material in a form convenient for practical applications. Considered are
the methods of experiments and the schemes of experimental devices;
principal rheological effects and dependences; the constitutive equations

xiv

Preface

consistent with thermodynamics, which are able to describe the whole set
of experimental data, and methods of evaluation of rheological parameters in these equations. Numerous comparisons are made between
predictions and experimental data that demonstrate which properties of
the constitutive equations are particularly related to one or another
effect. Applications of some results to industrial problems, some of them
of a new type, are also considered.
Much attention has been given to the experimental results observed in
highly nonlinear regions of uniaxial extensional deformation which have
been intensively studied in recent years. These include a growth of
effective elongation viscosity (as compared to a drastic decrease in shear
viscosity), the effects of elastic recovery and the regularities of relaxation
processes.
To describe the experimental data the constitutive equations used are
mainly those that have been obtained within the general approach 8,9
based on continuum mechanics, thermodynamics and division of complete strain into recoverable (elastic) and irreversible parts. Along with
the temperature, the elastic strain tensor is chosen as a state variable. The
stress, or its elastic part is then expressed in the same way as in the
problems considered for the crosslinked rubbers. Under discussion here is
a particular version of the general theory,8,9 the so-called nonlinear
Maxwell-like constitutive equations, which in the general case, contain
some unknown material functions. To specify these we proceed, from the
reasons of simplicity and agreement, between the calculations and data.
Godunov 1o independently proposed a specific nonlinear Maxwell
model which belongs to the general class 8,9 and has been successfully
used in solving some dynamic problems of the plasticity of metals.
Distinct from Godunov's model, the herein-described specification of
general Maxwell equations describes the flow-induced anisotropy in
viscous properties of polymeric systems and therefore results in qualitatively different effects. Although close in spirit, the particular model(s)
under consideration is still completely different from the highly specified
model proposed by Giesekus,1l,12 being much more justified and free of
some of the problems of the Giesekus model.
Using the constitutive equation of Maxwell type with flow-induced
anisotropy in viscous properties, the authors describe, at least semiquantitatively, the whole complex of viscoelastic effects obtained from
standard tests and well known for polydisperse polymer melts and
concentrated solutions of polymers; the accuracy of quantitative description commonly being 20-40%.

Preface

xv

The physical approaches to rheology are intensively developing nowadays (see, e.g. Refs 5 and 6). Unlike continuum mechanics, they can in
principle bridge the constitutive equations and molecular structure of
polymers or, at least, describe the parameters in these equations. But the
constitutive equations derived from microscopic models can be successfully used only if they do not contradict some general constraints
imposed by the laws of macroscopic thermodynamics. Also, the physical
approaches tend to describe the deformation of monodisperse polymers
which are fairly different in mechanical properties from the polydisperse
polymers used in industry. For instance, the elastic strains accumulated in
flows of monodisperse polymers are not too large whereas for polydisperse polymers they can be huge.
The first chapter briefly describes the class of general nonlinear
Maxwell-like constitutive equations used throughout the book, and then
introduces in more detail some useful specifications of the equations.
Additional information about continuum mechanics and irreversible
(non-equilibrium) thermodynamics, necessary for the comprehension of
this chapter, is given in the appendices.
The second chapter presents a review of simple constitutive equations
for viscoelastic liquids, based on the approach of continuum mechanics
and derived from molecular considerations.
The third chapter deals with analyses of general Maxwell-like and
single integral viscoelastic constitutive equations. Here a simple thermodynamic derivation of the Maxwell-like equations is given. Thermodynamics is also useful in analyses of Hadamard and dissipative
instabilities associated with the modelling of various terms in the equations.
The fourth chapter discusses experimental methods and schemes of
instruments used in polymer rheology. Particular attention is paid to the
extension of liquid polymers which is an unconventional test for the
common liquids.
Chapters 5-7 are concerned with theoretical and experimental studies
of the simplest regimes of deformation which are of interest for the
rheological characterization of polymeric liquids and throw light on the
specification of constitutive equations. The choice of the experimental
demonstrations was predetermined by the existence of corresponding
theoretical calculations.
Chapter 8 presents studies of viscoelastic flows in pipes and dies. Also
under discussion are some experiments and 2D numerical simulations of
entrance viscoelastic flOWS.13

xvi

Preface

Chapter 9 analyses non-isothermal regimes and particularly, the effect


of dissipative heating on the flow of polymers. Also under discussion is
the operation of a new type of sealing which is based on the Weissenberg
effect.
Chapter 10 presents studies of various viscoelastic fluid mechanical
problems which are of a certain practical interest. In particular, a new
method for certifying raw polymers is described which could be valuable
for industry. This chapter also demonstrates some nonlinear dynamic
effects in the 'bar' approximation. To illustrate these, the results of
numerical calculations are presented which describe the strike of a
viscoelastic jet of finite length against a rigid obstacle.
Chapter 11 presents a review of melt flow instabilities. Though more
attention is paid here to new experimental data, old (and seemingly
forgotten) important results are also exposed, along with some possible
methods for their theoretical interpretation.
Chapter 12 briefly reviews some experimental effects related to ruptures of polymeric liquids, their detachments from the wall, flow-induced
crystallization and other effects which so far have no clear theoretical
explanation.
The book provides a basis for subsequent calculations of flows of
polymeric liquids with high shear elasticity. In its design, it is intended for
researchers working in the field of polymer processing as well as in the
food and oil industries.

Acknowledgement

The authors are very thankful to friends and colleagues in Russia and the
USA for their generous support in the recent troublesome years and help
with everything. The authors are particularly indebted to the graduate
students of the University of Akron (USA): K.E.P. Adovale, K. Ebner, Y.
Kwon and M. Simbhabhatla, whose assistance made a valuable contribution to this book. The authors are also very grateful to the members of
their families for their encouraging support and patience.

CHAPTER 1

Constitutive Equations with a Recoverable


Strain Tensor as an Internal Parameter

1.1 INTRODUCTION
In very general terms, the peculiarities in flow of polymeric liquids can

be described as follows:
(i) In many cases, polymeric liquids can be considered as incompressible media since the ratio of pressure, p, to the compressibility
modulus, K, is almost always considerably less than unity
(K ~ l09Pa, 14. There are some exceptions however, related to
highly intense moulding processes, particularly injection moulding 15 where the effects of compressibility seem to be quite important.
(ii) In any flow of a polymeric liquid, the total shear strain contains
reversible (recoverable or high elastic) and irreversible components; in other words, in polymeric liquids, the elastic strains
always appear in flow.
(iii) Huge recoverable strains (as high as tens), sometimes much greater
than in crosslinked rubbers,16 can be observed in these liquids and
the reversible changes in sizes of specimens can vary greatly. In
these media, the development of flow can never be neglected
because the irreversible strains are commensurable with the reversible ones.

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Thus for a liquid polymer, the typical relaxation time, 8 = '11fl, where '1
is the viscosity and fl is the shear elastic modulus, is a fundamental
characteristic representative of the ratio of viscous to elastic forces acting
on a moving liquid particle. Associated with this typical relaxation time
is the Deborah number, De = 8VIL, where V and L are the characteristic
velocity and length of a flow. This dimensionless characteristic was
introduced in Refs 17-19 and plays a very important role, sometimes
similar to that of Re in Newtonian fluid mechanics, in analyses of
viscoelastic flows.
If a polymer system is chemically crosslinked, with the crosslinks
disposed at very few points in a chain, as in the case of vulcanized
rubbers, it makes the flow of the polymer impossible and still preserves
a capacity for the accumulation of large elastic strains. The equilibrium
theory of large elastic deformations of rubbers and rubber-like crosslinked systems is developed well enough 16 though with some reservations, as discussed in Section 2.2.2. This theory is known as the nonlinear
theory of elasticity with the elastic potentials specific to polymers. For
liquid polymers, however, the physical situation is much more complicated, mainly because of the nonlinear superposition of viscous flow and
large elastic strains.
The qualitative physics underlying flows of polymer melts by deformation of a temporary network structure might be pictured as follows. 9 The
entanglements of the molecules in a network are the result of secondary van
der Waals' bonding and hence may fluctuate both in time and space. For
long flexible polymer chains, the section between two adjacent entanglements contains many monomer units (or Kuhn's segments). When an
external force is applied, the entanglement fluctuations result in the
macroscopic flow of polymer liquid due to the ordered (on average)
displacement of the centres of gravity of the molecules. Meanwhile, the
network structure causes large elastic deformation, due to a specific
elasticity commonly attributed to the entropy variations of the long
sections of chains between the entanglements for time periods shorter than
the characteristic time period of entanglement fluctuations. The random
distribution of the entanglements in the fluctuating network means that the
distribution of chain length between entanglements, in terms of either
molecular mass or contour length, is also random. This produces a large
spectrum of fluctuation periods and relaxation times. The fluctuating
structure can be approximately considered to be made up of embedded
sub-networks; each sub-network having a characteristic length 1 and
characteristic time 8 1 . The networks are considered to interact weakly.

Constitutive Equations with a Recoverable Strain Tensor

This behaviour has been considered many times and various rheological equations have been derived from such a qualitative model (see
Chapter 2). It also helps in understanding the nature of elastic strains that
accumulate as a polymer flows, and how the strains relate to viscous flow.
When the polymer chains are very long and flexible, the elastic deformations can be considerable. Other phenomena in polymer flows, such as
the Weissenberg effect and highly nonlinear retardation effects in rotary
rheometers and under elongation, are very similar to the behaviour of
crosslinked rubbers when deformed, as has long been recognized. 16
Thus the most important feature of polymer liquids is that they can
accumulate large recoverable strains while flowing, which may predominate over or even suppress the flow itself. The suppression of flow is
possible due to the appearance of macroscopic regions of clusters that
have almost elastic properties, probably because of topological entanglements. This behaviour is seen in many effects, such as flow cessation, open
channel siphoning and fluidity loss, which are discussed in later chapters.
It follows that perhaps all nonlinear effects in polymer liquids may be due
to the accumulation of large recoverable strains; the strains becoming
larger, the larger the external forces or strain rates. Also, these strains
would induce anisotropy in such properties of the liquid as viscosity, heat
conductivity, and dynamic birefringence, as occurs in elastomers.
Thermodynamically, this behaviour would correspond to two possible
equilibria. The first is the true thermodynamic equilibrium at rest; the
second would be an incomplete (or 'frozen') quasi-equilibrium corresponding to the high elasticity of an elastomer and typified by large
elastic strains. This yields the seemingly paradoxical result that an
increase in shear will take the liquid from the true equilibrium through
non-equilibrium states to another quasi-equilibrium state. If this is so,
then the rheology of a polymer system, especially if its molecular weight
distribution (MWD) is narrow, can be described using a single, though
highly-nonlinear relaxation mode. Although such a description will not
agree with experimental data for low shear rates, it will become more
accurate as the shear is increased, provided that the loss of fluidity is
properly described. On the other hand, several relaxation modes will be
needed if the MWD of the polymer system is wide.
In this chapter, we consider first the nonlinear viscoelastic constitutive
equations for a single Maxwell model associated with a single relaxation
time. For any flow described by such a model, we can determine the
elastic (or recovery) strain in the course of unloading a small particle of
the medium, relieving it from the influence of adjacent ones. Under such

Nonlinear Viscoelastic Effects in Flows oj Polymer Melts

an unloading, the behaviour of the small particle in the medium is


assumed to be nonlinear elastic. In general, the Maxwell constitutive
equations obtained in such a way contain two unknown functions: the
free energy and the irreversible strain rate, which depends on the elastic
strain tensor and temperature. Any specification of these two functions
consistent with some fundamental restrictions, results in a specific viscoelastic model by which we can hope to describe the rheological
properties of real materials without violating the fundamental laws of
nature. Within these restrictions, the analogy with crosslinked rubbers,
considerations of simplicity and at least qualitative agreements with
experiments are used for evaluation of these functions in the context of
the deformation of polymer liquids.
As follows from the above physical picture, polymeric liquids are
characterized not by one but by several relaxation mechanisms with
relaxation times, Ok = 17k/11k> where '7k and 11k are the viscosity and
elasticity shear modulus in the kth relaxation mechanism. It is well
known 20 that above the glass transition temperature all values of Ok have
the same temperature dependence (usually Arrhenius-like), so the ratios
Ok/Oj = O(kj are numerical constants characterizing the relaxation spectrum. Therefore we can use the definition of the Deborah number even
for the multi-relaxation case related to the value of 0 averaged over the
relaxation spectrum.
A qualitative description of polymer flow assumes that each relaxation
mechanism is described by the nonlinear Maxwell model. It is also
assumed that these relaxation mechanisms have very different characteristic times and therefore, the interaction between these in the course of
deformation is negligible. These assumptions lead to the rheological
model which is usually called 'the set of Maxwellian elements connected
in parallel'. Here we do not go into the reasons for formation of the
relaxation mechanisms, which can be of a varying nature.
The very idea of involving elastic strains in the rheological description
of viscoelastic polymeric liquids with subsequent use of the theory of
nonlinear elasticity was seemingly first set forth by Weissenberg 21 and
Kuvshinsky.22 The first theoretical approach to the rheology of viscoelastic liquids with the separation of total strain into reversible and irreversible parts, and the subsequent use of nonlinear irreversible thermodynamics, was proposed in Ref. 23. However, only rheological models
with small (though still nonlinear) elastic deformations were derived
there. The general continuum mechanic and thermodynamic approach to
viscoelastic media, with the recovery strain tensor considered as an

Constitutive Equations with a Recoverable Strain Tensor

internal thermodynamic parameter was developed in Refs 8, 9, 24 and 25


on the basis of nonlinear irreversible thermodynamics and particular
viscoelastic kinematics. Various elements of the theory were also investigated in Refs 10 and 26-28. Its linear version had been developed
considerably earlier. 29 - 37 The theory of nonlinear Maxwell media was
proposed in Refs 8, 9 and 38--41 as a particular case of the more general
approach to nonlinear viscoelasticity. In these papers, certain specific
models within this approach are proposed. Finally it should be noted that
Oldroyd has independently proposed 42 almost the same ideas regarding
the thermodynamic approach to the rheology of viscoelastic liquids, as
the present authors are beginning to publish in the West.
Sections 1.2, 1.3 and 1.5 consider the basic ideas and formulations that
were used to develop the general theory of the Maxwell-like constitutive
equations under study. Then some specific models are discussed with
respect to the flow of polydisperse polymers.
1.2 DECOMPOSITION OF THE TOTAL STRAIN INTO
RECOVERABLE AND IRREVERSIBLE COMPONENTS
AND THEIR RELATION TO THE STRAIN RATE
At the end of the 1960s, some new ideas relating to the general principles
of derivation of constitutive equations began to appear in the literature.
They originated from Eckart's concept 43 of a 'variable relaxed state'
applicable to each point of a continuum. It was soon recognized that this
concept adds a new axiom to those used in rational mechanics and
thermodynamics 44 ,45 and does not relate to particular properties of
materials such as plasticity, viscoelasticity, etc.
The concept might be most easily explained when deriving a linear
viscoelastic equation, say, of a Maxwell type. Here one usually operates
with elastic, Ge , and inelastic, Gp , infinitesimal strains and further introduces the dependences of both the strains and their derivatives on the stress.
Hereafter the subscripts 'e' and 'p' mean 'elastic' and 'plastic', respectively,
and obviously originate from plasticity theory. Due to the evident
equality:
(1.1)

where G is actual total strain, it is possible, at least in principle, to


determine the inelastic strain Gp by 'instantaneous' element-by-element
unloading of the medium from stresses, and therefore measure the value

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

of ep Then the value of ee can be found from eqn (1.1). Because this
procedure can be performed at any instant t, it defines the 'variable
relaxed state' proposed by Eckart. 43
Shortly after Eckart, Kondo 46 independently introduced the same
concept of a 'stress-free natural configuration' and applied it to studies of
yield in crystalline materials. Later, Kondo's concept was reformulated
more precisely47-49 and several authors 5o ,51 working in the field of
plasticity independently employed it. The concept was applied to the
rheology of viscoelastic phenomena proposed in Refs 23 and 26. The
specific viscoelastic kinematics, following mainly the results published in
Ref. 9, are briefly developed below.
Along with the continuum law of motion, x = x(t, 1;), considered in
Appendix AI, we assume that a viscoelastic liquid at any time t, can also
be characterized by an unloaded state (11) which is defined by the release
of the medium from its stresses, element by element. Thus we have the
unloading relation: 11 = 11(t, 1;). Hereafter x and I; are Eulerian and
Lagrangian coordinates of a material point of the liquid. If we assume
additionally that the unloading relation is a locally-smooth one-to-one
dependence between the positions (11) and (I;) for almost all material
points of the continuum, then by using the local gradient relations:
(1.2)

we can introduce the tensors of gradients for the complete (</J), recoverable (</Je) and irreversible (</J p ) strains. Here, according to eqn (A.ll)
derived in Appendix 1, we employ the notation </J = F- 1 for the total
strain gradient. Equations (1.2) yield the relation
(1.3)

The factorial decomposition (1.3) of deformation gradient is, in fact, a


mathematical formulation of the principle of 'unloading in small' and
extends eqn (1.1) to large deformations.
We can also introduce positive symmetric Finger tensors for the total
and recoverable strains as follows:
(1.4)

where the superscript t denotes transposItIOn and, for the sake of


simplicity, we use hereafter the notation c for the Finger tensor instead of
its common notation C- 1. Usual volumetric considerations (see eqn
(A.32)) result in the equations:
det </J=po/p,

(1.5)

Constitutive Equations with a Recoverable Strain Tensor

where p, Po and Pp are the densities of the medium in the actual, initial
and unloaded states, respectively.
In much the same way as the velocity gradient, Vv, is represented
through the time derivative of tensor gradient (see eqns (A.39-A.41)),
differentiating eqns (1.2) and (1.3) with respect to time t under fixed ~ and
tt variables results in the kinematic relations:
(1.6a)

VV=VVe+VVp,
Vv t =

tb cp- 1,

VVtp =,,1,.
,,1,.-1. ,,1,.-1
'l'e .).
'l'p 'l'p
'l'e

(1.6b)

Here the overdots mean the time derivatives of the tensors under
consideration; v, Ve and vp are actual, reversible and irreversible velocities;
Vv, VVe and VVp are velocity gradients; e, ee and ep are respective strain
rates, symmetric parts of the velocity gradients; and w, We and wp are
respective vorticity tensors, antisymmetric parts of the velocity gradients.
In eqns (1.6), all the velocity gradients and their symmetric and antisymmetric parts are written in the Eulerian representations, i.e. in actual
'space' x. Using the same method as for the derivation of kinematic
relations (A.I-A.40) for total continuum, we can easily obtain from eqns
(1.6) the viscoelastic kinematic relation in question:
Ce -

Ce

VVe - (Vve)t Ce = 0

(1.7a)

By using the decompositions in eqns (1.6a), we can also write down the
kinematic equation (1.7a) in the identical form:
(1. 7b)
It was assumed in Refs 8, 9 and 28 that irreversible vorticity tensor wp
vanishes, i.e. wp=O. Thus

(1.8)
Certain reasons in favour of this important relation, which are outside the
kinematic considerations, are given below. Using eqn (1.8) reduces eqns
(1.7) to the final form:

t+cep +ep c=o

(1.9a)

where hereafter we omit the subscript 'e' and define the common
operations for upper-convected, t, and co-rotational or laumann, c time
derivatives as follows:

t=C-c Vv-(Vv)t c= c-ce-ec

(1.10)

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

If instead of the recoverable Finger strain tensor c = c-l, the Cauchy


tensor c- 1 = C is introduced, by multiplying eqn (1.9) from the right and
left by c -1 we can rewrite it in the identical form:
~-1+c-1ep+epc-1=O

(1.9b)

where ~ is the lower-convected time derivative defined as follows:


~=c+ce+ec

(1.12)

Finally we consider the volumetric relations that result from kinematic


equations (1.9). Multiplying eqn (1.9) scalarly by (2c) -1 results in the
equation:

"2 x dt x In(detc) + trep= tre

(1.13)

due to the volumetric relations (1.5), detc=(pp/p)2. Then using, additionally, the mass conservation relation written in the form:
d

- dt In(p/po)=tre

(1.14)

we can represent eqn (1.13) as follows:


d
d
- dt In(p/pp)+trep= -dt In(p/po)

whence
(1.15)
Until now, all the kinematic relations obtained in this section were
valid for the description of any continuum with elastic and inelastic
strains, including plasticity in metals and viscoelasticity in polymeric
liquids. The viscoelastic phenomena in polymeric liquids display, however, one particular feature which makes these materials very different
from the metals: after unloading, the polymeric liquids return to the initial
rest state with the same density Po (and temperature T); that is, of course, if
the isotropic pressure does not tremendously exceed the volume modulus
K. This physical reason immediately leads to the relation:
Pp=Po

(1.16)

Then eqn (1.15) with allowance for eqn (1.16) yields:


trep=O

(1.17)

Constitutive Equations with a Recoverable Strain Tensor

Equation (1.17) holds for elastic liquids regardless of whether they are
considered as compressible or not. When the volume effects are not
important, it is possible to use the incompressibility assumption which,
due to eqns (1.13H1.17), is written as follows:
tre= trep =0

detc= 1,

(1.18)

Instead of the Finger tensor c, we can also use other measures of


recoverable strain, which are some isotropic tensor functions of Finger
tensor. For instance, the Hencky strain measure h is introduced in the
usual way:
h= 1/2 Inc

In the vicinity of the rest state (c

(1.19)

15), the asymptotic relation


(1.20)

holds true due to eqn (1.19). Additionally, when the assumptions (eqn
1.18)) of incompressibility are employed, the relation trh=O also holds.
When Ihl ~ 1, eqns (1.9) and (1.20) result in the asymptotic kinematic
equation:
(1.21)
which is usually employed in linear viscoelasticity.
The difficulties and logical imperfections in the viscoelastic kinematics
developed above cannot be ignored. From the very beginning it was
understood that the components of the deformation gradient, 4>. and 4>p
in the decomposition (1.3) cannot be always compatible despite the
compatibility conditions being fulfilled for the total deformation gradient
4>. Thus Eckart developed his theory43 by using Riemannian geometry.
But it did not play any technical role because the kinematic equation (1.9)
which follows from Eckart's concept, operates in familiar Euclidean
space. Godunov 10 established a profound relation between the incompatibility and theory of dislocations.
The logical imperfections were employed several times in the formulation of Eckart's kinematics, when we used the non-kinematic concept of
stress in the definition of the unloaded state, as well as when arguing eqn
(1.8). Nonetheless, this approach results in easy and clear metric relations
for invariant characteristics of strains which will later enable us to
develop a class of viscoelastic constitutive equations of a differential type.
Also the advantage of the particular viscoelastic kinematics developed
is that they formulate the rheological equations in terms of recovery

10

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

strain, the tensor parameter that can, in principle, be measured independently of other rheological variables. Though these measurements are valuable in themselves in the characterization of polymers,
they also provide an opportunity to check independently the basic
assumptions of the theory. This is a favourable feature of this approach
as compared with other viscoelastic constitutive equations considered in
the next chapter.
1.3 DEPENDENCES OF STRESS ON ELASTIC (RECOVERABLE)
STRAIN AND TEMPERATURE

The elastic liquids under study are assumed to have no internal rotation
which means that the actual stress tensor is symmetric, i.e. (I = (It. In the
absence of internal rotations there is nothing to support a difference
between the vorticity tensors wand We' This is the main reason underlying
the use of eqn (1.8).
In all the papers that develop the approach under consideration, it is
assumed that a free energy function f (see eqn (A.35)) exists, which
depends on the elastic strain tensor c and temperature T, i.e. f = f(T, c).
The constitutive equations for elastic liquids are 'Maxwell-like' if the
actual stress (I is expressed via the recovery strain c with the same relation
as for elastic solids, i.e.
(1.22a)
Here, the tensor, (Ie defined on the right-hand side of eqn (1.22a), is called
the thermodynamic (or equilibrium) stress. Thus the Maxwell liquid can be
also defined as a medium where the actual stress tensor (I coincides with
the thermodynamic tensor (Ie'
Equation (1.22a) was derived by Murnaghan 52 who used a typical
thermodynamic argument, reasoning that there is no dissipation for
elastic solids. Equation (1.22a) being written in the equivalent form
(I/p

of/oh I

(1.22b)

states that the Hencky elastic strain hand a/p tensors are primary
thermodynamic variables related by true potential equality (1.22b). It
should be mentioned that the only physical reason to operate with eqn
(1.22a) in the thermodynamically non-equilibrium case of elastic liquids is
the assumption that polymeric liquids possess an 'elastic limit', which is a
quasi-equilibrium situation usually achieved on very rapid deformations,

Constitutive Equations with a Recoverable Strain Tensor

11

where the temporary entanglements in a macromolecular network act


like crosslinks in the crosslinked rubbers.
Equation (1.22a) generally holds for anisotropic solids. In the particular case of isotropy, when the free energy depends on the basic
invariants h of tensor c, i.e. I = I(T, 11> 12 , 13), eqn (1.22a) can be written
in the form:
(1.23)
Here Ik = allah and h are the basic invariants defined in eqn (A.26). In
an incompressible case, when P = Po = constant and 13= 1, we can introduce the elastic potential W(T, 1 1 , 1 2 ) = Pol defined with the accuracy of a
term depending on temperature, to obtain, instead of eqn (1.23), the
Finger formula:
(1.24)
Here the isotropic pressure is usually determined from boundary conditions, and
(1.25)

11 =trc,

are two independent basic invariants. Equation (1.24) is easily derived if


the modified elastic potential
(1.26)
is introduced. Here p plays the role of a Lagrange multiplier which
takes into account the incompressibility condition 13= 1. Then substituting eqn (1.26) into eqn (1.22a) results in eqn (1.24) where
p=p- W 2 oI 2
When elastic strains are small enough, eqn (1.24) transforms into the
linear Hookean relation:
(1=

-p(j

+ 2/18

(1.27)

where /1 is the shear elastic modulus and 8, the infinitesimal strain tensor
when h-+O).
Instead of free energy I, it is sometimes more useful to employ,
especially in non-isothermal situations close to adiabatic conditions, the
internal energy u depending on elastic strain c and entropy s. Because of
the thermodynamic relation (see eqn (A.60))
(8~h,

12

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

we can also rewrite the stress-strain relation (1.22b) in terms of internal


energy as follows:
(f

pou/oh Is

(1.22c)

It should be mentioned that thermodynamics impose some additional


restrictions on the possible functional forms of free energy j (or elastic
potential W). We know, for instance, that for small elastic deformations
when h'::::,8, where 8 is the infinitesimal Cauchy strain tensor, the free energy
is quadratic in 8 with two independent coefficients, the shear modulus p,(T)
and compressibility modulus K(T). Thermodynamic stability, i.e. the
minimum free energy in the rest state, leads to a definite form in which the
free energy increases monotonically and is concave in the vicinity of the
unloaded state. Hence it follows that p,(T) and K(T) are positive. It has been
founds-lo that this requirement of internal thermodynamic stability must
also be imposed on the deformed states of elastic and viscoelastic media
with arbitrary elastic strains. The general analysis is given in Chapter 3,
where the necessary and sufficient conditions of the thermodynamic
stability are discussed. These are formulated in terms of constraints
imposed on the first and second derivatives of free energy with respect to the
basic invariants. This analysis also shows that the violation of thermodynamic stability results in very high frequency and short-wave
Hadamard instability. We present here, only the necessary conditions for
the thermodynamic stability of the most important incompressible case:

(i= 1, 2,3)

where

Cj

(1.28)

represents the principal values of the recoverable tensor c.

1.4 BALANCE OF MECHANICAL ENERGY AND DISSIPATION


Kinematic equation (1.9) and the relation (1.22a) between stress and
elastic strain do not yet represent a closed set of equations. This is
because the tensor ep is still not expressed via other kinematic or dynamic
variables that can be measured at least in principle. Nevertheless, these
equations make it possible to determine the balance of mechanical energy
for the nonlinear, single-mode Maxwell-like model of elastic liquids.
Multiplying eqn (1.9) scalarly by the tensor poj/oc results in the
equation:
ptr(c oj/oc) + 2ptr(ep c oj/oc)-2ptr(ec oj/oc)=O

(1.29)

Constitutive Equations with a Recoverable Strain Tensor

13

Here the overdot still denotes the time derivative in the frame of reference
of a moving particle, and it is also easy to see that tr(O" Vv) = tr(O" e). The
first term in eqn (1.29) is represented as pj while the second and third as
tr(O"e p ) and tr(O"e), respectively. Then eqn (1.29) can be rewritten in the
form of mechanic energy balance:
(1.30)
The right-hand side of eqn (1.30) is the power (rate of work done) of
the stresses on the total deformations in an infinitesimal particle of
the liquid. The first term on the left-hand side of eqn (1.30) is the
elastic (storage) energy, while the second term, represents the power of
the stresses on the irreversible deformations of the medium. As shown
below, this second term is related to the mechanical dissipation (or lost
energy) in the system. In addition to eqn (1.30), there is also the balance of
kinetic energy, which is easy to obtain from the momentum balance
equation:
(1.31)
where the mass forces were neglected. Multiplying scalarly this equation
by v gives, after simple manipulation, the balance of kinetic energy (per
mass unit):
p' E = V' (v' 0') - tr(O" e)

(1.32)

Here the first term on the right-hand side is the divergence of the stress
flux in the continuum. Adding eqns (1.30) and (1.32) yields the total local
balance of the mechanic energy in the system under study:
(1.33a)
This equation can be also represented in the divergent form:

ata [p(f + E)] + tr(O"

ep) = V' Qm

Qm=v'O'-pv'(f +E)

(1.33b)

Here p(f + E) is the density per unit volume of total mechanical energy
in the system, which is the sum of kinetic and free elastic energies; Qm is
the total flux of mechanical energy including the convective (last) term in
the second equation (1.33b). These equations have the same form for the
incompressible case.

14

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

1.5 GENERAL DEPENDENCES OF THE IRREVERSIBLE STRAIN


RATE AND THE DISSIPATIVE FUNCTION ON THE ELASTIC
STRAIN AND TEMPERATURE
This section describes a procedure of definition in which tensor ep is
considered, making a closed set of constitutive equations (1.9), (1.17) and
(1.23) for the compressible case, and constitutive equations (1.9), (1.18)
and (1.24) for the incompressible case. It expresses the irreversible
traceless strain rate tensor, ep , in terms of the state variables, recovery
strain tensor, c, and temperature, T. The general representation for such a
dependence 8 ,9 is based on the approach of non-equilibrium thermodynamics 53 (see Appendix A2)).
Using the conservation laws and Gibbs' relation, the expression for the
entropy production, Ps, can be represented in the form:
(1.34)
with q being the thermal flux vector. Equation (1.34) is the famous
Clausius-Duhem expression. In accordance with the Second Law of
Thermodynamics, the right-hand side of eqn (1.34) is positive for all the
non-equilibrium processes and is equal to zero in the equilibrium.
Substituting the balance equation (1.30) into eqn (1.34) reduces the latter
to the form: 8 - 1o

T'P s = -T-1'q'VT+tr(u'ep )
(tr(u' ep ) = tr(u d e p ))

(1.35)

where u d is the deviator of the stress tensor and the second relation holds
due to eqn (1.17). Equation (1.35) clearly demonstrates that there are two
origins of entropy production (dissipation) in the Maxwell media: inhomogeneity in temperature distribution and irreversible flow in the
liquid which is the origin of mechanical dissipation, D=TPsl. It also
shows that the expression for the dissipation is represented in the
characteristic bilinear form LXk' Yk where X k represents thermodynamic
'forces' and Yk is conjugated to these thermodynamic 'fluxes'. The
approach used here proposes that some phenomenological constitutive
relations between these do exist, which are assumed to be quasilinear, i.e.
dependent on the state variables, T and c. From this assumption and the
considerations of tensor dimensionality, the general constitutive relations
for the medium under consideration have the form:
(1.36)

Constitutive Equations with a Recoverable Strain Tensor

15

(1.37)
where M is the rank-four mobility tensor and K is the second rank thermal
conductivity tensor, whose structures and properties of transposition have
been established in Refs 8 and 9. Equation (1.36) is the rheological
relation in question, whereas eqn (1.37) is the equation of heat conductivity. Both of them are generally anisotropic due to the flow-induced
anisotropy which occurs in flows of elastic liquids. Additionally, there is
the compressibility relation that can be written as follows:
(1.38)
This is the same condition that holds for the compressible elastic solids
since due to the kinematic constraint (1.17) in the pure Maxwell model
under study, there is no contribution of volumetric elastic forces to
dissipation. A more general situation is discussed below. Also, by using
the relation 13 = (PO/p)2, eqn (1.38) is represented in the form
(1.39a)
as in the case of a perfect gas. When, additionally, the density fluctuations
in the medium are supposed to be small enough, the linear compressibility relation
p=K(p-po)/Po

(1.39b)

can be used. It should be noted that the approximate expression (1.39b)


can be derived from exact formulae (1.38) and (1.39a) only under
condition: Il/K ~ 1. This condition is true for the rubber elasticity.
The thermodynamics also establish some constraints on the signs of
the kinetic tensors in eqns (1.36) and (1.37) as related to the Second Law.
Indeed substituting eqns (1.36) and (1.37) into dissipative equation (1.35)
reduces it to the quadratic form:
(1.40)

Due to the Second Law this form must be positive definite in nonequilibria. The two terms in this form are independent however, which results
in separate positive definiteness of both the kinetic tensors written in the
symbolic form:

K>O,

M>O

(1.41)

Though the dependence of K on c for the polymer melts and concentrated


solutions is unknown, for the solid polymers it has been found to be
important. 54

16

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The irreversible strain rate tensor ep depends on temperature T and the


recoverable strain tensor c. The latter means that ep is an isotropic tensor
function of tensor c subjected to constraint (1.17), i.e. traceless. Then in
the most important incompressible case when the stress tensor is represented by eqn (1.24), the general expression for ep is of the form: 8,9
ep = bl(c-tH d3)- b2(c- l -tH 2/3)

(1.42)

Here b l and b 2 are some scalar functions dependent on temperature T,


and basic invariants of tensor c, 1land 12 , are defined in turn, by eqn
(1.25). It is also possible to find the constraint imposed on the functions bk
by the Second Law. Substituting eqn (1.42) into the expression for
dissipation (the second two terms in eqn (1.34 yields:
D=T'PsIT=AblWl +Bb2W2+C(blW2+b2Wd

(1.43)

Here the coefficients A, Band C, represented as


A = 4/3(Ii - 31 2 ),

B=4/3(1~-1 d,

are positive definite if tensor c is positive definite. Thus subordination to


the Second Law demands that the bilinear form (1.43) of variables bk and
Wk is positive definite. If, in accordance with the necessary condition of
thermodynamic stability,28 we assume that W l and W 2 , as well as b l and
b2 , are positive (non-negative), the dissipative function will certainly be
positive (non-negative), too. Thus these are sufficient conditions for the
dissipative function to be positive definite. The question about the
thermodynamic stability for the elastic potential is still unclear because
the criteria of the stability involve higher derivatives of Wwith respect to
1 land fz.
We assume, additionally, that the general representation (1.42) for
irreversible strain rate tensor ep has a regular limit transition to the case
of linear viscoelasticity with small elastic deformations where:
ep~h/()

(c

-> ~)

(1.45)

where ()=1'//J1- is the relaxation time.


1.6 SPECIFICATIONS OF THE ELASTIC POTENTIAL
AND THE IRREVERSIBLE STRAIN RATE
Three scalar functions are to be specified in order to use the constitutive
equations for the single Maxwell model derived in the previous section.

Constitutive Equations with a Recoverable Strain Tensor

17

These are: the elastic potential Wand the functions b 1 and b2 in the
general expression (1.42) for irreversible strain rate e p Only the incompressible case is considered here.
First, we need to specify the elastic potential, W(T, Ii' 12)' In accordance with the qualitative physical picture of polymer liquids, the potential
should be chosen in the form considered for the crosslinked rubbers. But
even in this equilibrium case, no theoretical expressions for the free
energy have been derived, at present, that are valid over wide ranges of
variations in Ii and 12 , Thus one should use some empirical relations
known in the rubber high elasticity.16 When Ii and 12 are relatively
small, the Mooney-Rivlin potential,
W =(/l/2)[(1-a)(I 1 -3)+a(l2 -3)]

(/l, a>O)

(1.46)

is applicable, where a is a dimensionless, small positive fitting constant.


When a = 0, eqn (1.46) is reduced to the classical potential of the network
theory of rubber elasticity:
W = (/l/2)(I 1 - 3)

(1.47)

Equation (1.47) was derived from a statistical approach by using the


Gaussian assumption, which is not valid for very high strains. Therefore
the region of the applicability of eqn (1.47) to the real deformations in
crosslinked rubbers is very limited.
Other empirical formulae for elastic potentials have also been proposed
in the rubber-like elasticity. These are mostly expressed in terms of
principal values Ck of the Finger tensor c, which can in principle be
represented via the basic invariants Ii and 12 , though in a more
complicated way. Valanis and Landel 55 assumed that
(1.48)
where, from the symmetry considerations, the separate functions w(x) are
identical. The most important particular case of eqn (1.48) is the 'powerlike' relation proposed by Ogden 56 where
(1.49a)
Here Ilk represents the 'moduli' and ak represents some numerical fitting
constants. In all particular cases considered by Ogden, the products /lkak
were always positive. The important reason for this is discussed in the

18

Nonlinear Viscoelastic Effects in Flows oj Polymer Melts

next chapter. After substituting eqn (1.49a) into eqn (1.48) we can also
represent eqn (1.49a) in the following equivalent form:
W=

L 2J.1k (trc~k/2 k

3)

(1.49a)

iXk

The discussions of predictability of the general potential (1.48) and its


particular form can be found in the original papers cited and also in Ref.
16.

To the authors' knowledge, the widest range in description of various


experimental data for the crosslinked rubbers was achieved with the use
of the Blatz-Sharda-Tschoegl (BST) potential: 57
(1.50)
which includes the four non-negative constants: moduli J.1 and B, and
numerical fitting constants, nand m.
Later on we shall dwell on the simplified BST potential where B == 0: 38
(1.51)
which looks very similar to the single-mode (k= 1) Ogden potential (eqn
(1.49)) with the only distinction that in potential (1.51), parameter J.1 is the
Hookean modulus defined for the case of infinitesimal deformations
(where there is no dependence on parameter n), which is not the case for
the Ogden potential. Very good approximations of the experimental
results were demonstrated in Ref. 57 using potential (1.51) for simple
elongation with stretching ratios up to the order of ten. It should be
mentioned that hereafter, to simplify the notations, parameter n IS
introduced instead of nj2, as was done in the original paper. 38
In specifying the elasticity moduli, it is usually assumed that J.1 is
proportional to pT where p is the density and T is the Kelvin temperature. The same is supposed to be true for other moduli, J.1k, in model (1.49)
and B in model (1.50).
Now let us specify the dependence of the irreversible strain rate ep on
the recoverable strain tensor c. Apart from contraints (1.18) and the
Second Law requirements, there are no fundamental considerations
related to the principle of choice for the function ep No attempt has been
made to derive the set of constitutive equations under study from
molecular modelling and the reasons for this are discussed in the next
chapter. Thus the only means we have at hand to model the dissipative
term within the above constraints are the usual considerations of simplicity and fitting (at least qualitative) of the common rheometric data. In

Constitutive Equations with a Recoverable Strain Tensor

19

specifying this dependence, it is assumed 8 ,58 that the natural scalar


relation
(1.52)

is true. Here 1'1* is 'viscosity' which in general is a function of invariants 11


and 12 and sharply depends on temperature T. Equation (1.52) demonstrates isotropy in viscosity for viscoelastic liquids and was established in
Ref. 8 following the Ziegler variational principle. 59 Equation (1.52) was
also applied in Ref. 10 to plastic flows in metals, but it cannot be used for
describing the simplest properties of polymer melts such as normal and
shear stresses in simple shear (see Chapter 5). Thus we need to specify the
irreversible strain rate function by using the general anisotropic relation
(1.42). It was assumed 8 that as in the case of rubber-like solids, there is
such a relation between the material functions b 1 and b z which admits a
planar strain deformation for simple shear, where 11 = I z, and the
deformation in the axis orthogonal to the shear plane is absent. The
condition of existence of such a deformation is: 8 9 ,39
(1.53)
Further speculations about invariance: ep --+ - ep when c --+ C -1 and
- e, valid for elastic solids or for the total continuum, along with the
Ziegler principle, resulted in establishing the 'potential' relations:

e --+

for the scalar functions bk in eqn (1.42). Here, as in eqn (1.52), '1* is a
scalar viscosity, and W is a proper rubber-like elastic potential. Since for
real rubbers W(I bIz) # W(I z, 11 ), eqns (1.42) and (1.54) establish an
anisotropic relation between the tensors O'd and ep Equation (1.54) was
applied to describe the data within a multimodal approach in the case
'1* = constant and while this appeared to be good enough for shearing, it
proved to be unsatisfactory in the extension with large elastic strains 38
(see also Chapters 6 and 7). Although one more attempt 60 was made to
improve the situation, two lessons can be learnt from the comparisons: (i)
the anisotropy in the dissipative term is far more extensive than is predicted by the simple anti-symmetry property of rubber-like elastic potentials, and (ii) in good modelling, the properties of the elastic potential and
the dissipative term in kinetic equation (1.9) may not be related.
In Refs 8, 9, 25, 39, 40 and 65, eqn (1.42) was specified in the form:

b 1 =b z =e*,

ep =[c-c- 1 +o(Iz-I 1 )j3]/4e*

(1.55)

20

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

which is also a particular case of eqn (1.54) for the neo-Hookean rubber
potential (1.47). Here (J* is a scalar relaxation time which, in general, can
depend on temperature T and invariants Ik of elastic strain c. Of course,
there are many other possible ways of modelling the dissipative term
rejecting the considerations of planar deformations in the simple shear.
For instance, an easy relation for the tensor ep which is non-planar in
simple shear, can be written as follows:
(1.56)
This equation will be discussed in Chapters 5 and 6.
When (J* depends only on temperature, which is denoted as (J* = (J(T),
we can use the simplified Arrhenius dependence
(1.57)

(J=(Jo exp(EIRT)

with constant pre-exponent (Jo, activation energy E and gas constant R.


Equation (1.57) is valid if the temperature in elastic liquids is well above the
glass transition temperature. Equation (1.57) was found to be very
descriptive for polymeric liquids in a wide range of temperature variations
and has been tested by numerous experiments (see, e.g. Ref. 1). When
(J* = (J(T), eqns (1.55) and (1.56) satisfy the right limit transition to the region
of infinitesimal elasticity: ep -+ hl(J and substituting this limit in the linear
evolution equation (1.21) gives the linear equation of viscoelasticity. When
(J* depends on the invariants of elastic strain, this dependence has to be
normalized to provide the limit transition to the linear case.
Contrary dependences of the relaxation time (J* upon the elastic strain
invariants are proposed in Ref. 8. The first of them is related to the
orientation of macromolecules in strong mechanical (and other) fields and
results in the increase of (J* versus elastic strain invariants:
(J* = (J(T) exp[fJw(I1> 12 )]

(O~fJ~ 1)

(1.58)

where the constant Pis a 'hardening' parameter characterizing the rigidity


properties of the polymeric chains. If the value of parameter P is high
enough, the relaxation transition to the rubber-like elastic state can happen
in flows of liquid polymers and results in the onset of the fluidity loss
phenomena. These phenomena are discussed in more detail in Chapter 7.
On the contrary, the distortion of the potential barriers by stresses
proposed in activation theory of flow 61 62 cannot also be ignored. This is
taken into account by the Eyring dependence:
(J* = (J(T) exp( -

yaIRT),

(1.59a)

21

Constitutive Equations with a Recoverable Strain Tensor

generalized approximately for the 3D case. Here the constant y is


activation volume and (J is the stress intensity. Because Eyring's mechanism is effective only at large elastic strains when (J ~ W; W being an elastic
potential, eqn (1.59a) can be modified as follows:
O*~ O(T)

(1.59b)

exp( -yW/RT)

This relation was originally used in Refs 39 and 40.


It should be noted that the effects of mechanical degradation will also
exert a similar effect on the relaxation time. Which mechanism dominates in
flows of polymer systems is discussed in more detail in Chapter 7 and 12.
1.7 ON THE THERMODYNAMIC STABILITY AND POSITIVE
DEFINITENESS OF THE DISSIPATIVE FUNCTION FOR
THE MODEL SPECIFIED WITH EQNS (1.51) AND (1.55)
In the following modelling of polymer rheology, it shall be demonstrated
that the 'lucky' combination of the elastic potential (1.51) with the
relation (1.55) for the irreversible strain tensor possesses good predictive
ability in describing the whole set of data, at least for the polymer fluids
considered. In this case, one has to be sure that the necessary condition
(1.28) of thermodynamic stability is satisfied and the dissipation D defined
by the second formula in general formula (1.35) is positive definite. To
prove these properties it is easy to consider all the elastic terms represented in the principal axes of tensor c where the stress tensor (1 and
irreversible strain rate ep have diagonal matrices.
Let us derive first the expressions for W l and W 2 for the elastic
potential (1.51). Using the general relations:
W(I b 12) = (Jlj2n2)( C~ + C~ + C~ - 3)

where the incompressibility condition C1C2C2 = 1 was employed, and


differentiating these with respect to independent variables Cl and
C2(0 ~ Cb C2 ~ (0) gives:
W l == 8W= /1/(2n)
81 1 Cl-C2

(Cl

c~ -c~ -C2 c~ -C~)


Cl-C3

C2- C3

(1.60)

22

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Utilizing the expressions (1.60) in general Finger relation (1.24) is meaningless, and to calculate the principal stresses O"j, it is much easier to
operate with the formula:
O"j

= - p + (/1/n)c'J

(j= 1, 2, 3)

(1.61)

(i<j)

(1.62)

which gives:

Now the calculations using eqns (1.24) and (1.60) also result in eqn (1.61),
which proves the validity of eqns (1.60).
Straightforward and easy calculations with the use of eqns (1.60) show
that (1.28) ineqn are satisfied for any value of Ci if and only if
(Vi#j)

(1.63)

Inequality (1.63) when applied to the infinitesimal case Cj ~ 1 + 2h j shows


that the elastic modulus /1 is positive. Then, for any real value of
parameter n, and particularly, for any positive n, in eqn (1.63) is satisfied.
This means global satisfaction of the necessary condition (1.28) of
thermodynamic stability for the elastic potential (1.51).
In the principal axes of tensor c, the irreversible strain rate tensor ep
has the principal values:
(1.64)
When the principal values of stress and irreversible strain rate tensors
are represented by eqns (1.61) and (1.64), the dissipative function (1.35)
can be expressed as follows:
(1.65)
where
x=(1/n){(2c1-cz -C])[(Cl -cz)(l +C3)+(Cl -c3)(1 +C2)]
+(2cz -c] -cU[(cz -c3)(1 +Cl)+(C2 -cl)(l + C3)]

+ (2c] - c'i -CZ)[(C3 -cl)(l +C2)+(C3 -c2)(1 + cd]}


This expression for function X is represented in the form:
X= (3/n)[(1

+ C3)(Cl -

+ (1 + C2)(C3 -

C2)(C'i - cz)+ (1 + Cl)(CZ - C3)(CZ - c])

cd(c] - c'i)]

(1.66)

Constitutive Equations with a Recoverable Strain Tensor

23

Equation (1.65) and (1.66) demonstrate that due to the condition (1.63) of
thermodynamic stability, the dissipation D is positive definite if the
relaxation time e* is positive, whether the parameter n is positive or not.
The relation between thermodynamic stability (evolutionarity) and positive definiteness of dissipation was first discovered by Godunov. 1o
1.8 NONLINEAR MUL TIMODAL MAXWELL CONSTITUTIVE
EQUATIONS

The single Maxwell model considered above cannot describe such important properties observed in polymeric liquids as the spectrum of
relaxation times and strain recovery (retardation). Retardation is defined
as slow, inertia motion followed the instantaneous release of a viscoelastic
continuum of stresses. The retardation phenomena are described in more
detail in Chapter 4. There are two ways of including the retardation
phenomena in the theoretical approach developed. The first is to consider
various nonlinear relaxation mechanisms in the system which will also
describe the relaxation spectra observed for elastic liquids in the linear
region of deformations. The second way is to assume an inequality
between the actual and thermodynamic stresses. This section discusses
only the multimodal Maxwell approach related to the connection of
several (two as minimum) nonlinear Maxwell models in parallel,8,9,38.39.40.58 while the second possibility is briefly considered in
Section 1.11; details of both methods of theoretical modelling are developed in Refs 8 and 9;
Following the qualitative physical picture of imbedded slightly-interacted networks discussed in the introduction to this chapter, we can
propose the following thermodynamic relations:
(1.67)

Here fb Uk and Sk are free energy, internal energy and entropy for each
sub-network depending on a particular Finger recoverable tensor Ck'
Equations (1.67) yield
(1.68)

which shows that the usual thermodynamic equalities are satisfied. Now
we assume that for the kth sub-network the sub-stress O'k and irreversible

24

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

strain rate tensor ep,k are expressed by the same formulae as in the case of
the single Maxwell model. For the multimodal approach, the straightforward implementation of particular viscoelastic kinematics developed in
Section 1.1 is impossible and needs to be generalized. Nevertheless, one
can naturally define in this case, the thermodynamic flux associated with
thermodynamic force, Uk exactly by the same expression as was derived
from kinematic considerations. This approach is valid regardless of
whether the elastic liquid is considered as compressible or not. Also, the
multimodal approach has a transparent physical sense if, in accordance
with our physical picture, we assume that the relaxation times obey the
inequalities:

(k=I,2 ... )

(1.69)

Indeed, due to ineqn (1.69), every relaxation mechanism (or 'relaxator')


represents the whole liquid in a restricted time interval while the contributions of other mechanisms are approximately negligible. Within this approach, we can also consider both compressible and incompressible cases.
In addition to the thermodynamic relations (1.67) and (1.68), in the
multimodal Maxwell approach under study, the following dissipative
relations are valid:
TFs= -T-1q'VT+ ,LDk'
k

(1.70)

q=,LKk'VT=K'VT
k

Here Dk and Kk are the partial dissipation and the heat conductivity
tensor in the kth sub-network (or relaxator), respectively. It is usually
assumed that the heat conductivity tensor is isotropic and constant. In
the particular version of incompressible multimodal approach with the
specifications (1.51) and (1.55), the whole set of constitutive equations
takes the form:
(1.71)
(1. 72)
(1.73)

Constitutive Equations with a Recoverable Strain Tensor

tre=O,

detck= 1

25

(1.74)
(1.75)

11,k = trc~, 12 ,k = trck l(Jik = Jik,O TITo)

(1. 76)

where O~ are described by either eqn (1.58) or (1.59) with the constants 13k
or Yk and dimensionless elastic potential Wk or stress intensity O"k which
are related to the kth relaxator (or relaxation mode). The set of constitutive equations (1.71)-(1.76) is usually called the 'nonlinear Maxwell
equations connected in parallel' because in the linear limit, it relates to
the visual image of spring-and-dashpot models.
By virtue of ineqn (1.69), along with the complete set of equations
(1.71)-(1.76), one can also consider the so-called 'rapid adjustment'
approximation,38 when in terms of time scale t;::;;;Oko one can neglect the
fast relaxation processes beginning with k > ko and assume that the
irreversible strain rate is determined from the quasi-stationary relation
(see eqn (1.73)):
(1.77)
1.9 MUL TIMODAL APPROACH IN THE LIMIT CASE OF
LINEAR VISCOELASTICITY

When the intensity of flow in a viscoelastic liquid is very low, i.e. if Ie I ~


and the flow of liquid is near the true state of equilibrium, it is assumed
that there is a limit transition to the case of linear viscoelasticity. This
means that in eqns (1.71)-(1.76), the elastic sub-strains, Ck> can be
considered to be infinitesimal in limit (see eqn (1.27)):

(1. 78)
whereas the complete strains can be arbitrarily large and result in a
low-intensity flow. When the limit (1.78) holds, the set of equations
(1.71)-(1.76) is easily transformed into the following:
W=IWk,
k

Wk=Jik tr8~

(1.79)

0"= -p~+ LUb


k

Uk = 2Jik8k

(1.80)

a80at + ep,k = e,

ep,k =8k/Ok

(1.81)

26

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

tre=O,

Il,k

= 12 ,k = 4trsi

(1.82)
(1.83)

where 11k are the 'viscosities' in each relaxation element. Equations


(1.80) and (1.81) result in the differential form of Maxwellian linear
equations:
(1.84)
In 1D cases, the small deformations of viscoelastic liquid described by
the linear equations (1.84) are similar to those described by the springand-dash pot model connected in' parallel; the laws of the motion for the
dash pots and springs being similar to the Newtonian viscous and
Hookean elastic relations. The linear constitutive equation (1.84) and the
spring-and-dashpot model have the following correspondence:

1= Ip,o + 1.,0 + I.,k + Ip,k

(1.85)

where 10 = Ip,o + f.,o is the initial length of the model in the rest state and
1.,0 is the length of springs before deformation, f(t) is the actual length of
the model in the deformed state at instant t, and f.,k and 1p ,k are the length
increments in the kth spring and dashpot. This evident behaviour of the
mechanic model under loading develops good intuition in the comprehension of real deformations of viscoelastic media.
Integrating the first (differential) equation (1.84) with allowance for the
second, results in the integral representation:
(1=

-pl5+2

m(t) =

foo m(t-r)e(r) dr

L (11k/Ok) exp( -

t/O k )

(1.86)
(1.87)

Here the rest state is assumed at the time t -+ - 00. Equation (1.87)
demonstrates that the relaxation function, m(t), monotonically decreases
to zero and is concave downwards. Its structure also shows that only the
deformation history with time r ~ 0, exerts an appreciable effect on the

27

Constitutive Equations with a Recoverable Strain Tensor

stress at the instant t. A medium with such a feature is called a medium


with fading memory.

Also, the definitions of elastic potential Wand dissipation D given by


eqns (1.79) and (1.83), respectively, allow us to represent these quantities in
the form of quadratic memory functional of the strain rate tensor e('r) ('r ~ t):
W=

D= -

Itoo

(1.88)

m(2t- Tl -T2)tr[e(Tde(T2)] dTl dT2

Itoo

rh(2t- Tl -T2)tr[e(Tl)e(T2)] dTl dT2

(1.89)

Equations (1.86) and (1.87) can be also generalized by adding a pure


viscous term whose contribution to the stress is 21]oe; the fading memory
property being preserved with this additional term.
In formal derivations of the linear viscoelasticity63.64 for elastic liquids,
not only relaxation functions m(t) with exponential kernels (eqn (1.87)) are
used but those defined in the interval (0, (0) and having the properties:
rh(t) < 0,

m(t) >0,

Mi =

m(O) == Go < 00,

m(tO,

too tim(t) dt<

00

(i=O, 1,2),

M6/Ml

m(oo)=O,

M o =1],

(1.90)

=G e

Here Go is the 'instantaneous' elastic modulus, Ge is the high elastic or


'equilibrium' modulus, and 1] is the Newtonian viscosity. If an additional
pure viscous term is included, m(t) has a delta-Dirac component at t= +0
with a constant multiplier equal to 1]0' In this case, the complete
Newtonian viscosity, 1]=1]o+M o. The magnitudes 1] and Ge are most
often used for the rheological description of a polymer fluid. When the
relaxation function is represented by eqn (1.87),
(1.91)
Here 8 is the characteristic relaxation time evaluated by the relaxation
spectrum J-lk(8 k).

Using common integral transforms it is also possible to represent eqn


(1.86) in the inverse form written for the total strain Il as follows:
ll=ad/G O + 1i1

It

-00

ad(T)dT+

It

-00

m(t-T)daiT)

(1.92)

28

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where ad is the deviator of stress tensor and m(t) is the creep function with
the following properties:
m(t) >0,

dm/dt> 0,

m(O) =0,

(1.93)

There is also an integral relation between the relaxation and creep


functions which can be written in the form:
m(t)jG o +

J:

m(t-r)[1/11 +dm(r)/dr] dr= 1

(1.94)

A more detailed description of the linear viscoelasticity can be found,


for example, in Refs. 63 and 64.
1.10 ON THE RHEOLOGICAL PARAMETERS IN NONLINEAR
MULTIMODAL MAXWELL CONSTITUTIVE EQUATIONS
Every nonlinear Maxwell element in constitutive equations (1.71)-(1.76)
contains four independent positive rheological parameters. Two of them,
the relaxation time, (}k> and the elasticity modulus, Ilk> can be determined
from common rheological experiments in the region oflinear behaviour of
polymeric fluids where the linear equations (1.79H1.83) are supposed to be
valid. Thus the parameters (}k and Ilk can be found by the fitting procedure
from the dynamic experiments with low-amplitude shear oscillations, from
the stress relaxation tests and also from steady-state simple shear experiments. With little parametric modelling involving only a few Maxwell
models, only rough fitting of the data can be usually achieved, since
assumption (1.69) about a considerable separation of Maxwell models in
the multimodal approach is obviously rough. For example, the maximum
number of relaxators used in Ref. 38, in order to obtain a good description
of the data, was three, and these satisfied the following inequalities:
(1.95)

Examples of a more detailed description are given in Section 5.4. It is also


worth mentioning that a reasonably good description of the experimental
data in the nonlinear region of deformations is considerably better than
that in the linear one. Moreover, the variations in properties from batch
to batch for the industrial polymers, are so substantial that they almost
eliminate the tendency towards very accurate description.

Constitutive Equations with a Recoverable Strain Tensor

29

Thus with this kind of modelling, to predict the experimental data in a


wide region of deformations it is often quite enough to set the total
number N of Maxwell elements in the multimodal model equal to
2_3. 38 ,65 As a rule, there is no need to record a great number of
relaxators with a strong influence on each other in order to describe
quantitatively the flow of polymeric liquids in a limited region of
Deborah number variations. The same is also true for the high-speed
processing where the values of Deborah number are very large. But in the
latter case, rather than increasing the number of relaxators, more accurate modelling of the nonlinear elastic and dissipative terms is required,
taking into account some additional phenomena related to melt flow
instabilities (see Chapter 11) and dissipative heating. Sometimes, especially in the flows of polymeric liquids with a limited stress intensity, we can
simplify the rheological analysis of flows even for the small number of
relaxators (N = 3) involved in the modelling. Indeed, due to in eqn (1.95),
we can neglect the elasticity in the third relaxator and treat this as a pure
viscous element:
2

(/;::;:, -pt5+

(/k+ 2113 e

(1.96)

k=l

where 113 is the viscosity in the third relaxator.


The constants nk in the constitutive equations (1.71)-(1.76) as well as
the parameters 13k or Yk in relaxation times
being independent of
temperature, are responsible for the nonlinear modelling of rheological properties of viscoelastic fluids and therefore can be found only
from the experiments valid in an essentially nonlinear region of deformation.
Parameters nk could, in principle, be determined from the step-wise
high intensity deformations of the liquid because they describe the
'instantaneous' nonlinear elastic behaviour of polymeric liquids. Other
procedures are also possible. Sometimes even qualitative comparisons
between predictions of the model and data provide some constraints on
possible values of parameters nk' For instance, as shown in Chapter 5, the
comparisons for the simple shear give:

et

(1.97)
and, additionally, one can assume that all values of nk are the same, i.e.
There are also some methodical procedures for determining the
values of the constants from experimental data which will be discussed
later. As one example of these, we can mention that usually the process of

nk == n.

30

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

determination of constant n should start from the value n = 1 when the


parameters (}k and J1k have been found and the parameters y (or {3) are
taken equal to zero, with a subsequent increase in values of n.
Due to the principle of time-temperature superposition valid for the
polymers in the viscoelastic state (see Section 9.3), we can also usually use
a uniform value of activation energy E for all the relaxators (and
viscosities in the relaxators) assuming that Ek =' E.
There can also be some exceptional cases when a long-time relaxation
mode with a low elastic modulus is additionally introduced, for which
n1 #n and E1 olE (see, e.g. Chapter 6 and Section 7.4), but usually this
mode is not important.
Except for Chapters 7, 8 and 11, the hardening phenomena associated
with the increase in relaxation time, as shown in eqn (1.58), are ignored in
this publication. In order to use the Eyring effect (1.59) on relaxation time
one has to estimate the value of parameter y. Evaluations of the number
of segments involved in the activation process (see Ref. 1, Chapter 2,
Section 5; and Ref. 66, p. 378) result in the estimation: y~ 10- 20 cm 3
whereas in Ref. 39 the value of y was taken as ~ 10- 5 cm 3 . Though
Bartenev67 speculated that this discrepancy is due to 'structure variations
in polymeric liquids in the course of their deformation, in comparison
with solids', this difference is still so great that it is impossible to treat
the values of y in terms of Eyring's theory, rather than consider these
as empirical coefficients. Even with this large value of parameter
y~1O-5 cm 3 , this effect is negligible up to great values of elastic deformation.
1.11 ON THE SIMPLEST VERSION OF THE VISCOELASTIC
MODEL WITH RELAXATION AND RETARDATION
As discussed above, along with the modelling of viscoelastic properties
for polymer fluids by the multimodal Maxwell model connected in
parallel, there is also the possibility of describing more simply the
retardation phenomena in the liquids by introducing in addition to the
Maxwellian elements, a viscous term also connected in parallel. Let us
consider the simplest version of linear viscoelastic constitutive equations
which enable us to describe qualitatively both the retardation and
relaxation phenomena in viscoelastic liquids. These equations might be
obtained, in particular, from eqns (1.80) and (1.81) for the linear multimodal Maxwell model if the elastic strains are negligible in all Maxwell

31

Constitutive Equations with a Recoverable Strain Tensor

elements (k> 1) with the exception of the first one. Under this assumption, eqns (1.80) and (1.81) are reduced to the following:
08t!Ot+8t!(J1 =e,
ep,k ~ e,

(k= 1)

t11 =2J118b
t1k

~ 2IJke

(1.98)

where IJ is the Newtonian viscosity for the total model. Similarly to eqn
(1.84), eqns (1.98) can also be written in the simpler form:
(1.99)
(O~S=(Jr/(Jl

< 1)

Equation (1.99) has only three independent parameters: (Jb (Jr and IJ; (Jr
being called the retardation time. Because all of these parameters have
approximately the same temperature dependence as viscosity IJ, dimensionless parameter s has numerical values from the interval (0,1). Note
that unlike the Maxwell equations (1.84), in constitutive eqns (1.99) the
stress is not determined only by elastic strain. Equation (1.99) was
seemingly first introduced by Frohlich and Sack 68 to describe some
rheological properties of suspensions.
The derivations of linear viscoelastic constitutive eqns (1.98) and (1.99)
by using the approach of non-equilibrium thermodynamics can be found
in Refs 29-32 where, in particular, it is shown that the inequalities:
IJ > 0, (J 1 ~ (Jr ~ 0 follow from the Second Law of Thermodynamics.
The general theory of nonlinear viscoelastic models with relaxation
and retardation, which are reduced to eqn (1.99) in the linear limit, was
developed in Refs 8 and 9 by using the approach of quasilinear nonequilibrium thermodynamics. Some principles of duality between nonlinear extensions of linear models connected in parallel and in succession
have been established in detail in Ref. 9, where general theorems about
unloading and relaxation are proved and specifications of the model are
also given. At present, it is not clear which type of modelling has better
descriptive ability, the multimodal Maxwell approach considered in this
publication, or the nonlinear approach with linear limit (1.99), because
the latter has not been experimentally tested.

CHAPTER 2

Other Constitutive Equations for


Elastic Liquids

2.1 INTRODUCTION
Almost 40 years of extensive research of viscoelastic liquids have resulted
in a vast number of viscoelastic constitutive equations being brought on
to the rheology 'market'. There have been many attempts to derive them
from the viewpoints of mechanics, physics and mathematics, but no deep
understanding of the nature of viscoelasticity has been reached yet. This
is directly reflected by the fact that more than 10 popular rheological
equations are in competition nowadays. Some of them were derived on
the basis of a pure phenomenological, continuum mechanics approach,
some of them by using molecular arguments. Many of them can describe
some basic rheometric flows well enough and others not so well, even in
the very limited region of Deborah numbers. Yet, many of those with a
good descriptive ability of some viscometric flows, show very quick,
'explosive' instabilities in more complicated flows as has been demonstrated in numerical simulations since the early 1970s. Because experiments with polymeric fluids also demonstrate new kinds of instabilities
as compared with Newtonian fluid mechanics, there is a lot of speculation
in the literature about relations between the instabilities in viscoelastic
constitutive equations and those observed in flows of elastic polymeric
fluids. Thus the problem of choice of rheological equations for very
complicated industrial applications is becoming increasingly important.

Other Constitutive Equations for Elastic Liquids

33

In order to formulate some possible principles of choice of the equations


one needs to know, besides the descriptive ability, a brief historical
perspective of their development, along with formulations of some basic
restrictions imposed on different terms of constitutive equations. Also,
molecular considerations relating the rheological properties of polymeric
liquids to their internal structure seem to be very important. Therefore
this chapter, after a brief historical review of continuum mechanics, will
also discuss the most important (in the authors' opinion) molecular
approaches. A discussion of the origins of 'non-physical' instabilities in
viscoelastic constitutive equations is given in Chapter 3.

2.2 SOME POPULAR NONLINEAR VISCOELASTIC


CONSTITUTIVE EQUATIONS DERIVED FROM
CONTINUUM MECHANICS
As mentioned above, there is a great list of publications devoted to
derivations of viscoelastic constitutive equations from very different
viewpoints, and to give a review of the whole field here would make this
chapter too voluminous. Fortunately, Larson's recent book 6 serves as a
good review of the field of study and also contains some comparisons
between predictions of various constitutive equations and experimental
data. Thus we can restrict ourselves to only those remarks in the
literature that are relevant to the most popular constitutive equations.
The first nonlinear approach to the rheology of viscoelastic liquids was
pioneered by Oldroyd. 69 70 He postulated quasilinear and nonlinear
constitutive equations of differential and integral types related to the
external observable variables, the stress tensor (1, and strain rate tensor
e, and also elucidated some of the important principles of invariance.
Several other concepts were also developed by scientists such as Rivlin,
Green, Tobolsky, Ericksen, Lodge, and their numerous successors (for
reviews, see Refs 71-73). It was also recognized later that many of these
concepts were associated with the Oldroyd approach. A great many
rheological equations, both of differential and integral type, were produced and these were able to describe many of the properties of
viscoelastic liquids.
2.2.1 Constitutive Equations of Differential Type
We start with the simplest class of Maxwell-like differential nonlinear
constitutive equations for incompressible viscoelastic liquids, which

34

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

include the most popular rheological models of this type. These


can be described in the form:
8[t - e('t' e + e .'t') + IX .(t5 + 't'/,u)' tr('t' e)] + f('t') = 21/e

(2.1)

where 't' is the 'extra stress tensor' related to the actual stress tensor (1 as
follows:
(1= -pt5+'t'
(2.2)
e is the strain rate tensor subjected to incompressibility condition, tre = 0,
t is the co-rotational (or Jaumann) time derivative of tensor 't', defined by
eqn (1.10), f('t') is an isotropic tensor function, t5 is the unit tensor, p is
the pressure defined with the accuracy of an isotropic term, 1/ is the
viscosity, ,u is the elastic modulus, 8';;(,1//,u is a characteristic relaxation
time, ()( and e are numerical parameters ( -1 ~ e ~ 1). The first addend on
the left-hand side of eqn (2.1) represents the elastic term whereas the
second, f('t'), is the dissipative term. Equation (2.1) is an extension of the
linear Maxwell model and describes the nonlinear viscoelastic behaviour
of incompressible viscoelastic liquids. This equation is derived by using
the considerations of invariance, simplicity and fitting the data and, in
the particular cases discussed below, by using microscopic and thermodynamic arguments.
When:

()( = 0,

8 = const,

1/ = const,

f('t') = 't',

-1 ~ e ~ 1 (2.3)

eqn (1.1) represents the Gordon-Schowalter 74 or Johnson-Segalman 75


model which, in turn, includes the upper-convected (e = 1), lower-convected (e= -1) and co-rotational (e=O) Maxwell models. Apart from the
pure phenomenology, these models were derived by using some microscopic arguments (see e.g. Refs 5 and 6).
When:
()(=O,

-1 <e < 1,

f('t')='t"g(tr't'),

1/=constant,

8= constant
(2.4)

eqn (2.1) reduces to the Phan Thien/Tanner mode1 76 77 where the


function g was assumed to be linear or exponential. Equation (2.3) is a
particular case of eqn (2.4) when g == 1.
When:

()(=o,

f('t') = 't',

1//8 =,u = constant,

8=8(IIe),
lIe =2tre 2 ,
d8/d IIe~O,

eqn (2.1) represents the White and Metzner model. 78

(2.5)

35

Other Constitutive Equations for Elastic Liquids

When:
rt = 2/3,

11 = constant,

~=1,

e= J-l/11 = constant

(2.6)

eqn (2.1) reduces to the model independently proposed by Marucci 79 and


Larson 80 by using some microscopic arguments.
When:
rt=o,

(2.7)

~=1,

where f3 is a positive numerical constant (0 ~ f3 < 1), eqn (2.1) reduces to


the simplest Giesekus model ll 12 which describes phenomenologically
the orientation phenomena in elastic liquids.
When:
rt=o,

=1

and f(1:) has a complicated form which includes only one parameter, J-l,
eqn (2.1) reduces to the simplest Maxwell model 8 9 with neo-Hookean
potential (1.47) and kinematic equations (1.9) and (1.54, 1.55).
It should be also noted that all the equations summarized in eqns
(2.1) could be extended to the multimodal approach by representing
the extra-stress as the sum 1:=~1:K and inserting 1:k in eqn (2.1) instead
of 1:.
Equation (2.1) with a linear dependence f(1:) is a particular case of
Oldroyd's general eight-constant constitutive equation: 70

err - ~

(1: e + e1:) + rttr(1: e)t5 + f3etr1:] + 1:


= 211{e+ er[e- ~1 e2 + rtl tr(e 2 )t5]}

(2.8)

which is linear in extra stress 1: and at most quadratic in strain rate e.


Here e, er and 11 are relaxation and retardation times and Newtonian
viscosity, respectively, and ~, ~ b rt, rt 1 and f3 are dimensionless material
constants. Equation (2.8) also takes into account the nonlinear retardation phenomena by the term e omitted in eqn (2.1) for the sake of
simplicity.
If ~ # and rt = we can formally introduce, instead of extra-stress
tensor 1:, a new variable, the dimensionless symmetric second rank
configuration tensor, c, as follows:

(2.9)

36

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where it is supposed that c is positive definite. For the sake of


generality, we use in this chapter and in Chapter 3 (but nowhere else
in the book) the same notation for the tensor c as in Chapter 1, but
now it has no sense of recoverable Finger strain tensor C- 1 and even
in the incompressible case, generally detc;6 1. In Section 2.3, it is
shown how tensor c is expressed via microscopic properties of macromolecules.
Substituting eqn (2.9) into eqn (2.1) with a = 0 and ~;6 0 reduces eqn
(2.1) to the set:
[c-~(ce+ec)] +j(c, 0=0,

(2.10)

where the isotropic pressure is different from that defined in eqn (2.2) by
the term /lie. In this particular case, when ~= 1, eqns (2.10) are reduced
to the form:
(2.11)
where t is the upper-convected time derivative of tensor c, and for the
White and Metzner model (2.5), e and /l still depend on lIe.
It is assumed additionally that at the rest state, c=o. Then eqns
(2.10) and (2.11) will describe a continuous transition to the rest state,
if
lim f(c,

lim "'(c)-+O

~)-+O,

(c-+o)

(2.12)

and the functions f and g are regular functions of c.


Equations (2.11) have an obvious generalization:

t+ "'(c) =

(2.13)

0,

where (T.(c) is the 'equilibrium stress tensor' defined with the help of
elastic potential as in the case of elastic solids, i.e. by eqn (1.22). Some
features of these representations will be discussed in Chapter 3, but for
now it is enough to demonstrate how eqns (2.13) could represent the
general case for the class of Maxwell models with upper-convected time
derivatives, including the Larson equation (2.6) as a particular case. It
should be noted that Larson 8o has generalized eqn (2.6) and expressed
the extra-stress tensor (Te via configuration tensor c. He also found the
expression for the elastic potential W as follows:
(Te

= /lC/B(I1),

W = (3/l/2a) InB(I1)

B(J 1) = 1 + a(I1 - 3)/3,

J 1 = trc

(2.14a)

Other Constitutive Equations for Elastic Liquids

37

where IX is now a constant (0 ~ IX ~ 1). Substituting the expression for (1e


from eqns (2.14a) into eqns (2.1) and (2.6) and excluding after that the
scalar term [i 1 - 2tr(c' e)], results in eqn (2.13) where
t/I(c) = B(1 d(c- 15)/8

(2.14b)

General Maxwell-like constitutive equations (2.13) have been recently


derived from so-called Poisson-bracket formalism. 81 - 84 This approach
has been established previously for elastic solids 85 86 where the Hamilton
formalism of classical mechanics was extended to the case of continuum
mechanics employing functional approach and variation derivatives.
Grmela 81 82 was the first to extend this approach to the dissipative
viscoelastic case taking a dissipative functional into consideration. This
was done in the same way that the dissipative function is included in the
Lagrange method in classical mechanics. One can find the general
derivation of eqn (2.13) in Beris and Edwards' papers 83 84 along with a
detailed discussion of particular viscoelastic constitutive equations proposed in the literature. Being general and valuable for the analyses of
nonlinear stabilities, the Poisson-bracket formalism is awkward for
derivations of constitutive equations. That is why, for this purpose, one
of the present authors proposed a local generalized formalism of nonequilibrium thermodynamics. 41 This approach is discussed, along with
some comments, in Section 3.2 of Chapter 3.
Equations (2.10) could also be represented in a general form as eqns
(2.13) which seems to be the most general form for the viscoelastic
Maxwell model. But these equations are found to be generally unstable
with a very quick 'explosive' instability, seemingly with no physical
reason (see Section 3.3, Chapter 3).
2.2.2 Constitutive Equations of Integral Type
There is also a broad class of viscoelastic constitutive equations of
integral type, some of which were also initially proposed in Oldroyd's
paper. 69 It is possible to represent the stress-strain relation in the general
form of a superposition integral constitutive equation:
t'=

Loo

[4>1(11,1 2 ,

t - t1)C-1-<P2(1t.12, t-t 1)C]dt 1

(2.15)

which was first proposed by Rivlin and Sawyers. 87 Here C- 1 is the


Finger relative strain tensor for incompressible medium; 11 = trC- 1 ;
12 =trC; and <P1 and <P2 are generally independent functions. For the
description of viscoelastic liquids, eqn (2.15) is usually complemented by

38

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

the evolution equation (A.45) with the initial condition: C- 1 t=< = 15.
Unlike the constitutive equations of differential type, where the positive
definiteness of configuration tensor c should be proved, tensor C in eqn
(2.15) is always positive definite by its definition. This is an advantageous
feature of viscoelastic constitutive equations of the integral type.
When using the 'co-rotational' generalization of the Maxwell equation,
it is also possible to express the superposition integral in the form first
proposed by Goddard and Miller. 88 When using the Johnson-Segalman
differential approach, the constitutive equation can be also represented in
an integral form. 6 However, particular versions of eqn (2.15) are the most
popular today.
If the functions ({J1 and ({J2 are related by the potential expression:
1

au

({J1

=2 all'

(2.16)

where U is assumed to be a rubber-like elastic potential, eqn (2.15)


reduces to the Kaye-BKZ class 89 ,9o of integral constitutive equations.
Equations (2.15) and (2.16) were derived by using some thermodynamic
arguments and when step-wise deformations are applied they are reduced
to the relations of rubber-like elasticity. Despite a lot of differences, this
approach has much in common and is close in spirit to that analysed in
Chapter 1. A brief qualitative comparison between these two approaches
can be found on p. 160 of Ref. 6.
If:
(2.17)
eqns (2.15) and (2.16) are reduced to the Oldroyd-Lodge integral constitutive equation. 69, 72 When in eqn. (2.17) the relaxation function m(t) is
represented in the form shown in eqn (1.87), as a sum of decaying
exponents, this constitutive equation can also be represented as the
multimodal differential approach for the upper-convected Maxwell
model.
If:
(2.18)
eqn (2.15) represents the Wagner constitutive equation 91 ,92 which is a
good description of basic rheometric flows but bears no relation to
thermodynamics.
The predictive ability of the equations is demonstrated in Refs 91 and
92, and their mutual comparison is discussed in Ref. 6.

Other Constitutive Equations for Elastic Liquids

39

2.2.3 Rational Mechanical Approach

As mentioned above, the only relationship between the differential and integral viscoelastic constitutive equations can be established for specification
(2.17) of the general superposition integral constitutive equation (2.15). All
other constitutive equations represented either in differential or in integral
form seem to look very different. Therefore two fundamental questions
arise: whether or not a general theory unifying both approaches exists and
what general properties of viscoelastic liquids (and solids) are important for
such a theory. Such questions were resolved more than 30 years ago and
resulted in the creation of rational mechanics44 93 (also, for easy reading, see
Ref. 94). This approach includes particularly, the nonlinear theory of
viscoelasticity and is formulated in strict mathematical terms.
We start with some definitions of relative kinematics of arbitrary finite
deformations. Let a material point have the Euler's coordinates X t at the
time instant t, and X; at the time instant t'. Then the strain gradient
tensor Ft(t') is defined as follows:
(2.19)
The strain gradient tensor Ft(t') presents all the information about local
geometrical behaviour in the vicinity of a material point under study,
including not only strains but also the rotations of an elementary volume.
By using the Cayley polar theorem, it can be represented in the form:
(2.20)
where R is an orthogonal tensor describing the rotations of the elementary volume, and U is a symmetric (except for some 'monstrous' cases)
positive definite tensor that describes the deformations of the elementary
volume. The Cauchy tensor C is easily defined from the expression of the
square distance between two infinitesimally close points:
(2.21)
If the time shift s is defined as s = t - t', we can also write the tensor Ct(t')

in the following form


(2.22)
Such a tensor function of time for s ~ 0 is called the deformation history. If
compressibility effects are negligible, we can consider a medium as
incompressible with the constraint:
(2.23)

40

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Also, sometimes It IS more convenient to use, instead of the Cauchy


tensor Ct(s), the Cauchy deformation tensor,
(2.24)
An incompressible simple liquid is defined then, on the assumptions of
locality and causality that at a given material point, the extra-stress
tensor -r is represented as memory tensor functional % depending on the

deformation history Gls) at this point as follows:


s=

00

-r=%[Gt(s)]
s=o

(2.25)

The causality principle, well employed in many fields of physics, states


that in our attempts to describe natural phenomena, we are allowed
to know only the past, i.e. the history of a process but not its future.
This is reflected in the region of variations of the time shift s:
o::::;s< oo( - 00 < t'::::;t).
Additionally, it is assumed that the constitutive equations for the
simple liquid should satisfy the principle of material objectivity, which
means the invariance of constitutive equations relative to rigid (timedependent) rotations of the medium. Note that the momentum balance
equations do not satisfy this kind of invariance. Since the rotations of an
elementary volume in a medium subjected to deformations, are locally
independent of the strains, the principle of material objectivity means that
the strain gradient tensor Ft(t') is not valid for the descriptions of the
constitutive equations because it also describes the rotations. Thus one of
the kinematic variables (but, of course, not the only one) suitable for
constructing the constitutive equations for the simple liquid is the Cauchy
strain tensor Gt(s), as shown in eqn (2.25). It should be mentioned that the
principle of material objectivity was introduced at the very beginning of
this century by Zaremba95 who was also the first to understand the
importance of the principles of plasticity and viscoelasticity. Later the
same principles were independently re-established by Jaumann. 96 Good
illustrations in favour of these principles are given in Ref. 93 (p. 46) where
it is pointed out that the same Hook's relation is valid for a spring on a
rotating platform, independently of the speed of rotation, though both
the values for the stresses and strains are very dependent on the velocity
and acceleration of the rotation. It is also worth mentioning that though
in the special viscoelastic kinematics developed in Section 1.2, the
evolution equation (1.9) for the strain recovery tensor c satisfies the
principle of material objectivity automatically, the dependence of stress

41

Other Constitutive Equations for Elastic Liquids

on the pure deformation tensor c can be justified (as well as in the theory
of elasticity) only by the principle of material objectivity.
Until now, all the general formulations have been valid for both elastic
liquids and solids as well as for the plastic flows of such materials as
metals. The additional important feature which singles out the elastic
liquids as a particular class of materials, is the property of fading memory
which means that the deformation history in the remote past (s--+ (0)
contributes almost nothing to the stress at the present moment (s = 0).
This feature of elastic liquids has been used when postulating the fading
property of relaxation function met) at t--+oo, and the limit linear
viscoelasticity behaviour of the liquids has been considered (see Section
1.9). In order to introduce the concept of fading memory for the general
constitutive equation (2.25), it is necessary to choose a proper topology in
a functional space where the memory functional is defined. It is associated with estimating a difference between the values of the functional
for two 'close points' in the functional space. The choice of such a
topology could also define a characteristic relaxation time () in elastic
liquids. The conditions of smoothness chosen from the very beginning
and discussed in detail in Ref. 94 from the physical viewpoint, do not,
however, allow one to consider the pure viscous contribution to the stress
as well as the step-wise strain histories. Hence this formal approach was
recently extended to the non-smooth case 97 which results in far more
applications.
By using then a theorem by Frechet (see e.g. Ref. 98), when the strain
tensor Gt(s) is small enough, the general constitutive equation (2.25) can
be decomposed in the multiple integral form:
't"=

tXl m(s)Gls) ds+ tXl too {a(sb S2)G (Sl) "G (S2)
t

(2.26)
The multiple integral series (2.26) includes an infinite number of
unknown scalar characteristics of the material, 'relaxation functions' m, a,
p, ... with fading memory, i.e. decaying when s--+oo, Sk--+oo.
Also a general thermodynamic approach in the theory of simple liquids
has been developed,99.1oo which is popularly discussed in Ref. 94.
Though in the general case, the specific form of the functional (2.25) is
unknown and it is apparently impossible to determine it experimentally,
all the constitutive equations considered represent a particular form
specifying the general functional (2.25). For instance, in order to represent

42

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

eqns (2.13) in the form of expansion (2.25) we can use a procedure of


successive approximation for the solution of the first, differential equation
in (2.13) to find the evolution of configurational tensor c when the
difference (c- D) is small enough. Then substituting this asymptotic series
in the second, stress-strain relation in eqns (2.13) will give a particular
version of decomposition (2.25). As another example, we can consider the
general superposition principle expressed by the integral constitutive
equation (2.15). Under a simplifying assumption that the high-order
relaxation functions OC(Sb sz), /3(st. sz), ... are 'concentrated' on the diagonal Sl =Sz and their extra-diagonal contributions to the integrals are
negligible (i.e. oc, /3 are proportional to c5(sl -sz)), the functional (2.25) is
decomposed to a power series of the tensor Gls) to obtain after repeated
application of the Cayley-Hamilton identity the general expression in the
form of eqn (2.15).
It should be noted that the expansion (2.25) is not the only one. This is
not only because there are different tensor measures of deformations, but
also because we can expand the functional in terms of the Jaumann
tensor integral as shown in Ref. 88. Some other forms of integral
representations of differential viscoelastic constitutive equations written
in the form (2.8) are also possible. 6 ,z3
In spite of this uncertainty, the approach of rational mechanics is
sometimes useful in applications, especially in analyses of viscometric
flows where some general methods have been developed, and to some
extent, experimentally tested. 101 A constant deformation history serves as
an example. It is defined as follows:
Go(s)=exp( -ksNt)'exp( -ksN) (2.27)

where Q(Xt' t) is an orthogonal tensor and N(X t ) is a constant tensor and


k is a constant. The case NZO relates to steady-state shearing, whereas
the case N = Nt corresponds to steady-state stretching of the medium.
Both the flows are considered in more detail in Chapters 5 and 6.
Another useful application of the general approach (2.25) is the
derivation of constitutive equations for the 'second order liquid:'
-r=21'/0e+ /31ez + /3zAz,

A z = 2e+ 2(Vv)t.e+ 2e' Vv

(2.28)

where 1'/0 is the Newtonian viscosity, /31 and /3z are constants (/3z is
negative) expressed in terms of moments of linear relaxation spectrum
met), and A z is the second-order Rivlin-Ericksen tensor. Equation (2.28)
was derived 10z under the assumptions of 'short memory' and slowness of
the flow of viscoelastic liquid. The assumption of short memory means

Other Constitutive Equations for Elastic Liquids

43

that eqn (2.28) is valid only in a time scale that is much more than a
characteristic relaxation time e, i.e. for almost steady-state flows. Therefore the asymptotic equation (2.28) is non-applicable for solving a Cauchy
problem.
2.3 SOME PHYSICAL MODELS OF POLYMERIC LIQUIDS
All phenomenological approaches discussed above, however strict,
flexible and related to the thermodynamics they are, suffer from one
significant defect. The specifications of various terms in the constitutive equations, to some extent, are highly arbitrary and intuitive, though
sometimes they reflect important features of polymer structure, usually
found empirically. That is why, from the very beginning, a lot of
attempts were made to derive the constitutive equations for polymeric
solids and liquids on the basis of studying the deformation effects on
a molecular level, i.e. considering the motion of macromolecules and
also modelling some interactions between these. Such approaches are
very attractive and very popular nowadays, since they can in principle
provide us with an important relationship between the microscopic
structure of polymers and their macroscopic, e.g. mechanical properties. This section briefly reviews only those trends in the development of microscopic theories thought to be most important by the
authors. Readers with a greater interest can find more information in
the original references cited or in the recent reviews given in Refs 5
and 6.
2.3.1 Statistical Mechanics of Rubber Elasticity

We start with a briefreview of the theoretical treament of the equilibrium


case of rubber elasticity, which is traditionally considered as classic and
complete. To reduce the number of citations necessary for our demonstrative purpose, we will refer in this subsection mostly to standard
texts.16.103,104
The configuration statistics of a single chain in the statistical
mechanics of rubber elasticity operate with equilibrium distribution
function t/J o(R) of the end-to-end distance vector R, where in the Gaussian
case
(2.29)

44

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where b is the length of freely joined Kuhn's segment. Then the entropy of
a single chain is expressed as follows:
(2.30)

s(R) = k lm/J(R)
The force needed to keep this end-to-end distance is:
F(R) =

(2.31a)

Tos/oR

which for the Gaussian distribution (2.29) gives:


F(R) = 3kTR/b 2

(2.31b)

Equation (2.31b) shows that in the Gaussian chain, the force acts as in a
Hookean spring. Also very useful for applications, is the formula for the
macroscopic stress in a polymer (Ref. 5, Vol. II, Section 13.3):
O"=v fFRI/I(R)

(2.32)

dR=v(FR)

where v is the average number of chains per unit volume and hereafter the
symbol
is used for averages over the (generally non-equilibrium)
distribution function 1/1.
When considering a quasi-equilibrium, deformed, state of the macromolecular chains in a crosslinked rubber, the formula for distribution
function I/I(R) is changed; for example, for the Gaussian statistics, it is

<>

( 3)3/2 [3

exp - 2b 2 tr(C RR) (detC)1 /2

I/I(R)= 2rcb 2

(2.33)

Here C is the Cauchy tensor. Equations (2.29) and (2.33) are usually
derived from the random-walk considerations. Equation (2.33) can also
be derived by using equilibrium distribution function (2.29) and the
evolution (Liuville) equation for distribution function (Ref. 5, Vol. II,
p. 361).
(2.34)
under the affinity assumption:
(2.35)
where Vv is the macroscopic velocity gradient tensor. If the common
formula

c5S = k dRI/I(R) In[I/I(R)N(R o )]

(2.36)

Other Constitutive Equations for Elastic Liquids

45

for the deformation entropy of the whole continuum is used, where t/!(R)
is expressed by eqn (2.33), it results in the well-known expression:
[)S = - k/2(trC- 1 - 3)

(2.37)

which serves as a starting point for the classical statistical theory of


rubber-like elasticity. This theory was extended for the case of nonGaussian statistics by using the inverse Langevin distribution function
and also employing simplified FENE approximation (Ref. 5, Vol. II,
Section 13.5).
Though this theory is highly respected and often cited, from the very
beginning it was surrounded by mystery. For instance, the successive
development of the theory could not predict the Mooney equation (1.46)
which is quite descriptive in a medium range of elastic deformations in
rubbers and can be easily derived by continuum mechanics. A surprising
feature of this classical theory is that the Boltzmann's entropy (2.30) is
inconsistent with the definition of Gibbs' entropy employed in statistical
thermodynamics:

S[t/! J = - k dRt/!(R) lnt/!(R)

(2.38)

This feature has been recently discussed 105 and when using the Gaussian
distribution function (2.33) it leads to the surprising result:
[)S=S[t/!] -S[t/!oJ =0

(2.39)

meaning that for a Gaussian chain, there is no contribution at all to


the Gibbs' entropy! This was a starting point for the criticism of
classical theory in Ref. 105. The spring-and-bead statistical model
developed in Ref. 105 confirmed that the Gibbs' entropy variations
in the model appear only when the anharmonic terms are included in
the interaction of beads. Also, surprisingly enough, for the harmonic
bead interaction, expression (1.47) for the classical elastic potential was
confirmed in this paper; however, no Gibbs' entropy variations were
apparent.
2.3.2 Statistical Mechanics of Dilute Polymer Solutions
Historically, this branch of irreversible statistical mechanics of polymers
was first studied in a qualitative paper by Kargin and Slonimsky before
the Second W orld War. The results were explained in more detail later by
the same authorsl06.107 and also by Gotlib and Volkenstein. 108 The

46

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

approach was subsequently developed by Kirkwood and Riseman,


Rouse, Zimm and Cerf,109-112 among others. Only dilute polymer
solutions in low molecular solvent are considered in this theory, with so
small a concentration of polymer that interactions between very separated macromolecules are neglected. Thus two kinds of interactions are
important in this theory: intramolecular ones and the interactions between a macromolecule and solvent; the latter being mainly of a hydrodynamic type. A macromolecule is modelled by a set of beads
connected by springs where all the masses of macromolecular (Kuhn's)
segments and hydrodynamic interactions are concentrated on beads
whereas in the springs model the intramolecular, 'elastic' interactions are
between adjacent segments. Usually, Gaussian chains are considered with
the Hookean spring forces described by eqn (2.31b). Sometimes the
'internal viscosity'108,112 is also included in the intramolecular interactive
forces which models the potential barriers in internal rotations of
monomer units in chains relative to each other. 103 In the simplest
considerations, the hydrodynamic forces are modelled by the familiar
Stokes relation which is proportional to the difference in the velocities of
a bead and solvent. Additionally, the Brownian forces are taken into
account in the interaction between a macromolecule and solvent. They
are related to the 'bombardment' of each segment by small molecules of a
solvent. This very simplified modelling was simplified even more by
Bird. 5 In order to clarify the qualitative behaviour of macromolecules in
dilute solution, Bird considered instead of the chain macromolecule, the
'dumb-bell' i,e. an aggregate consisting of only two beads connected by
the spring. The force balance of a dumb-bell (Langevin equation) is:
mti+((u- Vv t R)+ F(R) =A

(2.40)

where u is the velocity of dumb-bell centre of mass, v is macroscopic


velocity of the whole continuum and , is the drag coefficient. Equation
(2.40) includes inertia, drag, spring F(R), and Brownian force A. This is a
stochastic equation and when F is a linear function of R, as in the case of
the Hookean dumb-bell (eqn (2.31b)), the properties of stochastic force A
are well defined and represented by the Gaussian (i-correlated stochastic
process. Otherwise some complications appear, as discussed below. It
should be noted that , is a phenomenological drag coefficient whose
uncertainty is related to the uncertain value of the bead radius, and Vv is
the macroscopic velocity gradient.
In the case of the Hookean dumb-bell, by using the intermediate
asymptotic procedure developed by Chandrasekhar,113 the Langevin

47

Other Constitutive Equations for Elastic Liquids

equation (2.40) is reduced to the governing Fokker-Planck equation for


the non-equilibrium distribution function, which combines the Liuville
equation (2.34) with the asymptotic expression for the probability flux, R,
expressed as follows:

R = (Vv' R - F Is)t/I - (2kTI,)at/l loR

(2.41)

Here the usual extension on a non-Gaussian spring was made, though


without any justification. Substituting eqn (2.41) into eqn (2.34) yields:

Vi + alaR' [(V v . R -FIOt/I -(2kTlOat/llaR] =0

(2.42)

This is the Fokker-Planck equation for the non-equilibrium distribution


function t/I, which has to be solved under the natural initial condition:
t/ll 0 = t/I o Equation (2.42) describes the convective diffusion in the
configuration space {R}. The macroscopic term Vv considered asymptoti(=

cally as being independent of microscopic coordinates of the configuration


space, is assumed to be related here to the velocity gradient of the whole

solution, but not the solvent. This assumption, accepted as self-evident in


the literature, relates this kind of derivation to the so-called 'selfconsistent field approximation'. Though eqn (2.42) is linear, it can be
solved in a closed form only in very exceptional cases, which creates
many difficulties in the calculations of statistically averaged values.
Sometimes, when it is possible to find its solution explicitly, in the same
cases, the expression for the second correlations <RR) are also found
explicitly (Ref. 6, Section 6.4). Consider, for example, the Hookean
dumb-bell. The stress tensor in the dilute solution of such dumb-bells, due
to eqns (2.31b) and (2.32), is represented as follows:
t1

= Il<RR),

Il = 3vkT Ib 2

(2.43)

where Il is the elastic Hookean modulus. Also, substituting eqn (2.31b)


into the Fokker-Planck equation (2.42), multiplying this equation by the
diadic RR and then integrating over whole configuration space, results
after integrating by parts in the following evolution equation for the
averaged diadic <RR):
d/dt<RR) - <RR)' Vv - (Vv)t. <RR) + RR) -

b)le = 0

(2.44)

where e= ,b 2 j(12kT). When comparing eqns (2.43) and (2.44) with eqns
(2.11) where l/J(C)=C-b, which represents the upper-convected Maxwellian model via configuration tensor c, one can find their identity if
c= <RR). This is why the tensor c is called the 'configuration tensor'.
Moreover, because of this representation, the tensor c should be evidently

48

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

positive definite. Thus only the solutions of eqn (2.44) that preserve this
positive definiteness have physical sense. We shall discuss this matter for
the general case in more detail in Section 3.3.5 of Chapter 3.
These results are easily extended for the more realistic, bead-and-spring
Rouse model (see e.g. Ref. 6, Section 2.7) where the same transformation
to the normal modes is possible as in the completely linear case, described
in earlier papers.106-108,110,111 This transformation to the normal relaxation modes results exactly in the multimodal approach discussed in
Section 2.2 with the same elastic moduli for each relaxator and the
familiar Rouse relaxation spectrum. Also, the pure viscous contribution
of solvent to the stress can be incorporated into the modelling, but
usually it is done on the level of phenomenology. Some recent attempts
were also made to use more strict derivations from first principles,
including the solvent contribution to the stress (see, e.g. Ref. 114 and the
bibliography contained therein).
The parameters of the macroscopic constitutive equation (2.11) as well
as of its multimodal generalization are not completely described in
molecular terms, since the value of drag coefficient , is still uncertain.
Nevertheless, this way of derivation undoubtedly means progress in our
understanding of viscoelastic phenomena because it establishes an important relation between macroscopic and microscopic parameters in the
model.
In spite of this success, a lot of problems remained unsolved in the
more realistic modelling of the viscoelastic behaviour of dilute polymer
solutions. Some of these problems were discussed by de Gennes (Ref. 115,
Chapter 6) and in Ref. 6, Chapter 8. First, the long-term hydrodynamic
interaction being taken into account even in linear Oseen approximations,lll essentially changes the structure of normal modes. In a more
detailed approach, however, accounting for the long-term hydrodynamic
interactions leads to a nonlinearity in the basic Langevin equation.
Second, the assumption of the spring being Hookean is generally not true
for good solvents even in weak viscoelastic flows, not to mention the
strong ones. Though the corrections for non-ideal elastic behaviour can
be easily made for the dumb-bell model, this supplies the second kind of
nonlinearity to the basic Langevin equation. Third, the Rouse theory
(and all nonlinear possibe improvements mentioned as well) completely
neglect the effect of the possible formation of topological knots and loops
which can be only justified for good solventsYs Fourth, even in the
dumb-bell approach, all the nonlinear terms mentioned make it impossible to use the basic assumption of the 'c:5-correlation' of Brownian forces

Other Constitutive Equations for Elastic Liquids

49

in the stochastic Langevin equation (2.40). In this case, Chandrasekhar's


procedure for deriving the evolution equation (2.42) for the distribution
function from the Langevin equation (2.40), is not more valid, along with
the representation of Brownian force in eqn (2.42). This is the most
fundamental difficulty and some measures to overcome it are discussed
in Refs 116 and 117. Fifth, the appearance of nonlinear terms in eqn
(2.42) results in the 'statistical nonlinearity' of this equation, even if
the Brownian terms are completely neglected. This means that in the
equation for the second-order correlations, <RR), the correlations of
higher orders also appear. It leads formally to the infinite set of
non-closed equations. All attempts to overcome this difficulty by making some closure approximations are highly arbitrary and sometimes
very doubtful.
2.3.3 Microscopic Models of Polymer Melts and
Concentrated Polymer Solutions
Green and Tobolsky118 were the first to propose a nonlinear viscoelastic model for flows of molten polymers, based on the concept of
temporary network discussed in detail in the introduction to the
Chapter 1 (Section 1.1). This theory considered only one temporary
network formed by sections of macromolecules between the entanglements, with the same averaged contour length and the same 'lifetime'
for all the entanglements. Then, easy considerations of transition probability applied to the temporary Gaussian network resulted in constitutive eqns (2.15)-(2.17), with relaxation function met) represented by a
single exponent. This model was then extended on the multiple relaxation
times by Lodge 119 ,12o that corresponded to the concept of 'embedded'
temporary non-interacted networks. These results were later summarized
in Ref. 72.
When the relaxation spectrum is discrete, all these results can be also
obtained by using the Fokker-Planck equation (2.42), so it was natural to
find out a new physical interpretation for the derivations demonstrated in
Section 2.3.2 as applied to the molten polymers and concentrated
polymer solutions. Thus the results obtained in the previous section for
the dilute polymer solutions and interpreted in terms of dumb-bell
statistics, were re-interpreted in terms of statistics of strands. 6 The
concept of strand was then employed to a great extent in the microscopic
derivations of nonlinear viscoelastic equations for the concentrated
polymer systems. Though a definition of 'strand' is not given in Ref. 6, it
is understood here to mean a large section of a macromolecule between

50

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

temporary entanglements, which contains a constant number of monomer units (or Kuhn's segments). With this interpretation, many results
obtained for the dilute solutions are hoped to be also valid for molten
polymers. But the style of physical description is getting more phenomenological. For instance, in order to derive the constitutive equations
(2.1)~(2.3) from eqn (2.42), one has to change the term Vv t in this equation
for [Vv t -(1- e)e] which was proposed by Gordon and Showalter on the
reason of 'slippage' of the strand in the medium (see aso Refs 76, 77 and 6,
Section 5.2). In order to describe the effects of finite extensibility of the
strands, the FENE approximation for the dependence F(R) was employed
(Ref. 5, Vol. II, Section 13.5). There the statistical nonlinearity was
overcome with such an assumption as transposition of averages in
nonlinear expressions: <F(x) ~ F( < x . Anisotropic effects were also
included phenomenologically in the kinetic approach (eqn (2.42)) for both
the drag and Brownian term (Ref. 5, Vol. II, Section 13.7), which are
supposed to describe the flow-induced anisotropy in the elastic liquids.
The same method of transposition of averages was employed for decoupling correlations.
Though all these assumptions are not very different from pure phenomenology, the most questionable, in the authors' opinion, is using in
the strand approach, the same concept of Brownian interaction of a
strand with surroundings as used in the theory of dilute polymer
solutions. Indeed, in contrast to the dilute polymer solution, every strand
in a polymer melt is surrounded by similar strands and interacts with
these very differently from that described by the Brownian term in eqn
(2.41).
In 1956, Yamamoto121-123 developed another kinetic approach con-

sidering the breakage and creations of entanglements in the temporary


network as irreversible stochastic processes. In the simplest case of
Gaussian chains when expression (2.43) for the stress holds, his kinetic
equation is written as follows:

/I + a/oR' (Vvt RtjI) = g(R)- h(R)tjI

(2.45)

where g is the probability rate of entanglement creation and htjl is the


probability rate of entanglement destruction. Yamamoto also proposed
that g and h are functions only of R. Phan-Thien and Tanner 7 6 also
extended this equation allowing 'slippage of the strands' in the same
manner as discussed above. Kinetic equation (2.45) is free from the
above-mentioned defect related to the Brownian forces and follows, at
least qualitatively, the concept of temporary network. When approaching

Other Constitutive Equations for Elastic Liquids

51

equilibrium, I/I(t, R)--+l/Io(R) and the right-hand side of eqn (2.45) should
go to zero which yields:
g(R) = h(R)l/Io(R)

(2.46)

If in eqns (2.45) and (2.46), h= constant = 1/8, where 8 is the relaxation


time, multiplying eqn (2.45) by RR and integrating over the whole
configuration space results once again in eqn (2.44), and in turn, in the

upper-convected Maxwell model. But the physical sense of the solution is


now different from that described by the kinetic model (2.42) as well as
the solution of kinetic equation (2.45).
If the average amount of entanglements is conserved, the relation
J= f(g-hl/l)dR=O

(2.47)

holds true. Otherwise when J #- 0, integrating eqn (2.45) over the whole
space yields a kinetic equation for the variation of entanglements. In
general, both the functions g and h in eqn (2.45) have to be determined
from more profound physical studies. Even when the linear Yamamoto's
estimation (2.46) of the right-hand side of eqn (2.44) is valid, the same
statistical nonlinearity appears due to the dependence h(R). Also, some
pre-averaging procedures were proposed in the literature to obtain
viscoelastic constitutive equations (Ref. 6, Section 6.3). The activity in this
field was characterized in Ref. 6 (p. 175) as follows:
... The level of empirism involved in ... all the network models, is so
high that one should not take their molecular or structural derivations
too seriously. These models ought to stand or fall by their predictions
of stress, not by the strength of their molecular underpinning.
Serious progress in the understanding of the complicated processes
happening in flows of polymer melts, was achieved in a series of Doi
and Edwards papers.124-126 In these papers, Doi and Edwards developed
the reptation mechanism for motion of macromolecules in polymer
melts, proposed earlier by de Gennes. 127 This mechanism reflects the
simple idea that every macromolecule in a melt suffers from the constraints created by its neighbours. These constraints create an effective
'tube' along which the macromolecule only can move. The contour
of the tube is chosen stochastically by the 'head' of macromolecule,
with all the macromolecule's 'body' simply repeating the motions of
the head as in the motions of reptiles. De Gennes also developed a

52

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

dynamics of reptation based on a diffusion equation and obtained, for


the fraction of tube remaining unvacated, the following time dependence:
cp(t) =

(S/n2k2) exp(k 2 t/8 d )

(2.4S)

1,3, ..

where 8d is a characteristic time of reptation. It was also found that


8d '" M3 where M is the molecular mass of macromolecule, though in
contrast to the familiar dependence 8d '" M 3 . 4 (see, e.g. Refs 124-126).
Doi and Edwards completed this work. They considered first a
continuum deformation due to the independent reptation motions of
macromolecules in various 'tubes' and found the deformation to be
expressed by a specific strain tensor:
(2.49)

<

>u
where u is the unit vector along a particular tube and the symbol
means averaging over the distribution function for the orientation of the
unit vector u. Second, Doi and Edwards used simplified statistics similar
to those discussed for the network theories to obtain the final stressstrain relation:
n=v(M/Me)kT

foo cp(t--r)Q{F- (r)} d-r


1

(2.50)

where v isthe number of macromolecules in the unit volume, Me is the


molecular mass of the macromolecule between two entanglements, and cp
is the time derivative of the function cp(t) defined by eqn (2.4S). Equation
(2.50) falls into category (2.15) of viscoelastic constitutive equations of
integral type.
It is worth mentioning that the function Q{ F-1(t)} has no explicit
expression and needs to be calculated for each type of flow. This
inconvenience stimulated many attempts (Ref. 6, Section 4.5) to find a
simplified constitutive equation of differential type which could save all
the basic features of the reptation approach and its basic predictions.
Some of these attempts 79,BO have been mentioned when discussing eqns
(2.1) and (2.6). Also, Larson BO attempted to extend the Doi-Edwards
theory by taking into account partial extension of the tube, ignored in
their original approach. His constitutive equations are represented by
formulae (2.13) and (2.14). Despite this being an empirical approach, it
improves the appreciable discrepancies between the predictions of the
reptation theory and experiments.

Other Constitutive Equations for Elastic Liquids

53

The reptation theories, as formulated, have taken into account only the
slowest relaxation process. Their possible extensions to the region of
rapid relaxation processes are considered in Refs 128-130.
The reptation models cannot describe microscopic features of flow in
polydisperse polymers where long chains are moved in the low molecular
environments. Neither can they be appropriate for polymers, even monodisperse, with extremely high molecular mass, where too few macromolecular ends are presented there to organize the 'tube'. It should also
be mentioned that Curtiss and Bird131.132 developed independently, a
more strict approach to the physics of polymer melts with almost the
same final results as compared to eqns (2.49)-(2.50).
Finally, we briefly discuss an independent statistical approach employed in Refs 133-137 to derive some multimodal viscoelastic constitutive equations. This approach, which starts with detailed Brownian
dynamics of the bead-and-spring model in the low and high molecular
environments and for dilute polymer solutions, is reminiscent of the
Rouse-Zimm theory. When applied to polymer melts, this approach uses
a self-consistent field approximation which considers the statistic behaviour of a single macromolecule in a viscoelastic liquid. In this case, the
Brownian forces are derived from non-Markovian statistics and the selfconsistency condition relates the maximum relaxation time of the environmental liquid to the relaxation time of the whole continuum. Only
slow and 'super-slow' relaxation processes in Gaussian approximation
were considered. A remarkable feature of this approach is the introduction of small- and large-scale macromolecular motions which, after an
averaging procedure, result in the interaction of the two relaxation
processes on the macroscopic level. A good agreement between the
predictions of the approach and data is found when the Deborah
numbers are not too high.
In conclusion, it should be noted, that a very remarkable feature of all
the differential models derived from any molecular approach is operating
only with tensor e and, in more nonlinear equations, with only its first
invariant, 11 = tre, but never with tensor e -lor other invariants of tensor
e, 12 and 1 3 , This reflects the difficulties with the statistical nonlinearity
and closure approximations made to uncouple the correlations.

CHAPTER 3

Analyses of Simple Constitutive Equations for


Viscoelastic Liquids

3.1 INTRODUCTION
To construct viscoelastic constitutive equations properly one needs to
know, apart from their predictability, how robust they are and what
possible underwater reefs one may meet that could ruin a weak rheological ship in the course of its difficult route through the ocean of solving
complicated problems.
This is the main reason for some possible justification of the constitutive equations in their relation to the fundamental macroscopic laws of
thermodynamics. Another field where thermodynamics seems to be
helpful is in evaluating a physical sense of instabilities that have been
mathematically established.
This chapter mainly analyses the general Maxwell-like constitutive
equations (2.10) and (2.13). First, it is demonstrated that the general form
for eqns (2.13), derived from the Poisson-bracket approach,81-84 is easier
to obtain by using a local approach of irreversible thermodynamics for
both compressible and incompressible cases. Also, the derivation is
generalized to take models (2.10) into consideration. Then some problems
of stability for these constitutive equations are studied as associated with
the modelling of elastic terms (Hadamard stability) and with the
modelling of dissipative terms (dissipative stability). Here, the analysis of
Hadamard-type stability for the single integral constitutive equations

Simple Constitutive Equations for Viscoelastic Liquids

55

(2.15) is also briefly discussed. It is shown that the configuration tensor c


is positive definite for any piecewise smooth strain history. Simple
sufficient conditions are found for the boundedness of stress tensor at
given strain history. The results are then applied for some popular
viscoelastic constitutive equations.
3.2 THERMODYNAMIC DERIVATION OF GENERAL
MAXWELL-LIKE CONSTITUTIVE EQUATIONS
We will employ in this section, the general approach of irreversible
thermodynamics which has been discussed in Sections 1.3-1.5 and in
Appendix A2 of this book. It is assumed that in the simple case under
study, the state variables of elastic liquids are: the temperature T and the
symmetric second-rank configuration tensor c which is considered as a
typical internal parameter. Additionally, it is assumed that the tensor c is
positive definite and an attempt is made to justify this assumption later.
Being mostly interested in studies of isothermal flows of isotropic
elastic liquids, we can now introduce as a proper thermodynamic
potential the free energy density per mass unit,f=f(T,/ 1,/ 2,/ 3), where
Ik represents the basic invariant of tensor c. Then the 'thermodynamic
stress tensor' associated with free energy f, is represented as in the
quasi-equilibrium elastic case, again by eqn (1.23) and we can only rewrite
this as follows:
(fe = 2pc' ofjoc=2p[flC+ f2(I1 C-C 2)+ f31 35]
(II = trc,
12 = 1j2m - trc 2),
13 = detc)

(3.1)

It should also be noted that no assumption of incompressibility has been

made and the incompressible case will be considered later. Considering


further the isothermal behaviour of elastic liquids and using the dissipative relation (1.35), we can rewrite this as:
D= TPsl T =tr((foe)-pjIT,

pjlT =tr((fe 1j2c- 1 oc)

(3.2)

The next step in developing the approach is to represent the dissipation


as a characteristic bilinear form D = LXk Yk where X k represents thermodynamic 'forces' and Yk represents conjugated thermodynamic 'fluxes'.
As thermodynamic forces, we can naturally take the actual, (f, and
equilibrium, (fe, stresses. Then the thermodynamic flux conjugated to
actual stress (f is certainly the strain rate tensor e. Another thermodynamic flux, ee, conjugated to the quantity (fe is a tensor that is

56

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

related but by no means is equal to 1/2c - 1 "C, because the latter even does
not satisfy the frame invariance conditions. This demonstrates an ambiguity in the definition of the second thermodynamic flux, ee, associated
with the arbitrariness of the problem: how to find a tensor from a scalar
product of two tensors; one tensor being known. Thus an additional
physical assumption is needed to define the quantity ee' This has been
proposed in Ref. 8 as follows: not only thermodynamic force, lTe, but also
conjugated to this, thermodynamic flux ee, has to be defined as in the case
of quasi-equilibrium elastic solids. This assumption seems to match
reasonably well the local equilibrium hypothesis. It also results in the
following procedure: to find the solution, e, from the kinematic relation
for the co-rotational or Jaumann time derivative of tensor c, c=c"e+e"c,
valid for elastic solids (or for the total continuum) and hold the expression for the non-equilibrium case. This is exactly the same procedure as
that used in the definition of the equilibrium stress tensor lT e In so doing
we define the quantity ee searched for, as the solution of the equation:
(3.3)
The quantity ee can be expressed in an explicit form, which is too
awkward to reproduce here (see eqn (3.22) in Ref. 8). By using eqn (3.3)
it is easy to see that the term pj Iy in eqn (3.2) is reduced to the tr(lT" ee).
For the Maxwell liquids where IT=lTe, we can represent the expression
(3.2) for the dissipation in the form:
(3.4)
Finally, following the procedure of non-equilibrium thermodynamics, we
should establish a quasilinear phenomenological relation between the
thermodynamic force lTe and the generalized thermodynamic flux ep This
relation is of the form:
(3.5)
Here M is a rank-four mobility isotropic tensor function depending on
configuration tensor c, which is positive definite and symmetric in the
first two and second two indices and in transposition of the first two and
second two indices as well. Note that eqn (3.5) is reminiscent of eqn (1.36).
Further on, substituting eqn (3.5) into eqns (3.3) or (3.4) results in the
evolution equation for the configuration tensor c:

1: + t/I(c) = 0,
1: =C-C" Vv-(Vv)t.c,

t/I(c)=2c"e p (c)
t/lij(C) = 2CikM kjst(C)u e,st(c)

(3.6a)

57

Simple Constitutive Equations for Viscoelastic Liquids

which coincides with that presented in eqn (2.13). Here t is the upperconvected time derivative and t/I(T, c) is an isotropic tensor function of
tensor c defined as shown in eqn (3.6a) and related to the dissipative
processes in the elastic liquid. Additionally, in the rest state when
c---+~(Il---+12---+3, 13 ---+1), we assume that t/I(c)---+O. We also assume that this
limit transition, described by eqn (2.12), exists and is regular, meaning
that there is a limit to the linear viscoelastic Maxwell equation when the
intensity of the strain rate is very low.
Multiplying the first eqn (3.6a) from both the right and left by c- 1 ,
represents this equation in the 'dual' form:
(3.6b)
where ~-1 is the lower-convected time derivative of tensor c- 1 For the
incompressible case with linear dependences of (Ie and tfr(c 1 ) on tensor
c-t, eqns (3.1) and (3.6b) are easily reduced to the usual form of
lower-convected Maxwell model. The difference between these lower- and
upper-convected models is mainly in the dependence of stress on elastic
strain. When (Ie = I1C it corresponds to the case of rubber elasticity, while
the case (Ie = -I1C- 1 can describe the elasticity of crystals or metals.
If the new kinetic rank-four tensor L (c) is introduced
L ijk1

Cim C jn

114mnkl

then the dissipative term t/I(c) in eqn (3.6a) can be rewritten as:
(3.7)
Substituting eqn (3.7) into eqn (3.6) represents the latter in the form which
coincides with the evolution equation obtained by Beris and Edwards
(see eqn (2.13) in Ref. 84). The structure of tensor L has been revealed in
Ref. 84 and of tensor M, in Ref. 9.
In order to take eqns (2.10) into consideration also, we can now generalize formally the thermodynamic derivation of Maxwell-like constitutive
equations. For this reason, instead of natural, quasi-equilibrium relations,
(I=(le and eqn (3.3), we postulate the following non-equilibrium relations:
t/lij(c) =

(I = (leg,

Lijk1PO!/OCkl

~=e(c'e+e'c)

(3.8)

Here, the difference (e -1) characterizes non-equilibrium properties of the


model. In the spirit of eqns (2.1) and (2.10), we assume that
is a
numerical parameter (e E [ -1, 1J). Then the same procedure leads to
eqns (2.10) with the stress tensor, now defined by the first expression in
eqns (3.8), and dissipation being equal to D = tr((I'f(c, e)).

58

Nonlinear Viscoelastic Effects in Flows oj Polymer Melts

Thus it is shown that to obtain the equations for the general Maxwelllike model (and also other rheological models for isotropic or anisotropic
liquids) it is, at least, not necessary to use the Poisson-bracket formalism.
The definitions of thermodynamic forces and fluxes made in the natural
way and the Clausius-Duhem inequality, is all one needs for the derivation. The reason for this simplicity lies in the fact that the fundamental
expression (3.2) for dissipation is local and therefore one does not need to
employ variational procedures to derive the constitutive equations. But
the variation formulation can be very helpful in the analyses of nonlinear
instabilities (see, e.g. Refs. 85 and 86). Also, it has been demonstrated that
the local formalism of irreversible thermodynamics is not necessarily
related to the constraint trep = assumed in Refs 8 and 9 and employed
in Section 1.2. Some arguments in favour of this constraint are discussed
in Section 3.5.1.
In order to consider further the common incompressible case where
tre == V . v = 0, let us first analyse the volumetric condition in the general
quasi-equilibrium (~=o) Maxwell model (eqns (3.1) and (3.6)). Multiplying scalarly eqn (3.6) by (2C)-1 and using the mass conservation eqn (1.14)
yields

(3.9)

Equation (3.9) shows that


tre p =0,

p~po

when

c~~. It

also shows 8,9 that when

(trep =0)

(3.10)

exactly as in the equilibrium limit of elastic solids.


Thus the evolution equation for tensor c also describes the law of
mass conservation in a compressible elastic liquid only if the condition
trep =0 is satisfied. When tre p #0, the density variations are not described
more by the variations of the configuration tensor c but, in addition to
eqn (1.14) for mass conservation, satisfy the kinetic equation (3.9). This
means that in this case, density is not included as a thermodynamic
variable. An attempt to improve the situation by making the assumption
that f = f(T, p, c) is not successful because it contradicts the transition to
the quasi-equilibrium elastic limit. This strange situation was the only
reason why in the authors' publications,8,9 as well as in the model
proposed by Godunov/ o the condition (1.17), trep = 0, was additionally
employed.

Simple Constitutive Equations for Viscoelastic Liquids

59

Consider now the incompressible case when P = Po = constant. When


trep #0, eqn (3.6) still holds whereas eqn (3.10) is reduced to
tre = 0,

t d/dt(ln 13 ) + trep =

(3.11)

Surprisingly enough, in this case, the free energy remains the same, i.e.
f=f(T,1 1 ,1 2 ,1 3 ), because the tensor c is not directly related to the
density. The actual stress (I is represented by:
(3.12)

Here the isotropic term is introduced to satisfy the condition of incompressibility tre = and the thermodynamic tensor (leis still defined by eqn
(3.1). This means that the tensor (Ie is not defined with the accuracy of an
isotropic term as in the equilibrium case, and therefore the isotropic
pressure is in non-equilibrium. These features result in the fact that even
in the incompressible case, expression (3.4) for the dissipation D is now
not invariant under the transformation (Ie --+(le + p~.
For the incompressible case with ~ # 1, the evolution (first) eqn (2.10)
still holds, i.e. expressed as follows:
[~-~(ce+ec)] + f(c, ~)=O

(-1~~~1)

(3.l3a)

but the stress tensor and dissipation are defined as:


D = tr[(le f(c, ~). c- 1 ]/(20

(3.l3b)

In the particular case of the neo-Hookean potential (eqn (1.47)), the


constitutive equations (3.13) are reduced to eqn (2.10).
This analysis thus shows that only under condition tre p = 0, when
c = Ce- 1, the dependence of stress on the configuration tensor described
by the same relations as in the equilibrium case of elastic solids, otherwise
the stress tensor, even in the Maxwell case under study, is in nonequilibrium.
3.3 SOME INSTABILITIES IN VISCOELASTIC MAXWELL-LIKE
AND SINGLE INTEGRAL CONSTITUTIVE EQUATIONS
The arbitrariness involved in continuum mechanic derivations, even
when provided with some elementary thermodynamic restrictions, as well
as in microscopic derivations which also include a lot of empirism, could
result in the occurrence of the various instabilities that have been
reported in many numerical studies since the early 1970s. Many of these

60

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

instabilities have no physical sense and are related to poor modelling of


either elastic or dissipative terms in constitutive equations.
3.3.1 Hadamard Instability: Method of 'Frozen Coefficients'
First, we consider the possible instabilities related to the modelling of
elastic terms in general Maxwell-like constitutive equations (3.13), i.e.
the terms in the square brackets in eqn (3.13a) and the expression for
stress in eqn (3.13b). These are usually related to the so-called 'Hadamard
instability' (or 'ill-posedness', or 'non-evolutionarity') of constitutive
equations.
We refer to the complete set of equations as evolutionary (or
Hadamard stable) if the transient Cauchy problem for the set has at any
time, t, the solution which provides the complete initial conditions at
subsequent time moments. 13B This definition only means that contrary to
ill-posed problems, the solutions of well-posed problems can be continuously prolonged from the past to the future at any time instant t.
The importance of the evolutionary behaviour for the rheology of
elastic liquids was first discovered by Rutkevich139-142 and, to some
extent, by Godunov. 1o Some significant results were obtained by Dupret
and Marchal 13B and Joseph and coworkers. 97 Usually the Hadamard
stability is analysed by studying the behaviour of a characteristic cone
(see, e.g. Refs 97 and 138). This general method is, however, cumbersome.
In the case of quasilinear equations, i.e. linear in derivatives, one can
employ without loss of generality, the method of 'frozen coefficients'.41,97
The method is related to easy local stability analysis for the simple
infinitesimal, extremely short and high frequency waves of disturbances
propagating with a finite speed. Later in this section, we will apply the
method to the analysis of Hadamard stability of constitutive equations
for incompressible elastic liquids. We first analyse the stability of Maxwell-like constitutive equations (3.13), and at the end of this section, we
consider the Hadamard-type stability for the single integral constitutive
equations.
For the general Maxwell-like liquid under study, the total set of
equations consists of constitutive equations (3.13), the condition of
incompressibility and the equation of momentum balance. Substituting
the expression for stress in eqns (3.13b) into the momentum balance
equation allows us to represent the latter, along with incompressibility
condition, in the form:
Po (VI +v'Vv)= -Vp+R:Vc,

V'V=o

(3.14)

Simple Constitutive Equations for Viscoelastic Liquids

61

Here
_ 1 (O(J e,ij
-",-

R ijn1 - 2J;
..

UCnl

o(Je,ij )
+ -",-

(3.15)

UCln

It is assumed here that -1 ~ ~ ~ 1(~ #0), Additionally, we use the


evolution equation (3,13a) written in the form:
ct +V' Vc-1/2(1

+ ~)[c' Vv + (Vv)t 'c]


+ 1/2(1-~)[c'(Vv)t + Vv'c] + f(c, ~)=O

(3.16)

Equations (3.14H3.16) are written in a Cartesian coordinate system {x}


where (VV)ij =OVj/OXi' These equations represent a closed set relative to
variables c, v and p, which are linear in derivatives. It is also assumed that
this set of equations has a 'basic' solution {c, v, p} which satisfies some
proper initial and boundary conditions. Consider now some disturbances
{bc, bv,bp} of the basic solution. So the new possible solution with the
disturbances changes now for the following:
Cl

=c+bc,

VI =V+bV,Pl =p+b p

(3.17)

If we substitute eqns (3.17) into eqns (3.14H3.16), taking into account


that the basic solution, {c, V, p}, satisfies the latter equations, and the
initial and boundary conditions, we obtain some nonlinear equations
for disturbances with coefficients depending on space and time variables,
and satisfying zero initial and boundary conditions. We can also
simplify the problem considering the disturbances as extremely small
or 'infinitesimal' and linearize the equations for disturbances relative
to their small values. This procedure relates to the problem of linear
stability which is still very difficult to analyse because the coefficients
in the linearized equations for the disturbances (the so-called 'equations
in variations') are still dependent on the time and spaces variables.
Consider now a particular case when the disturbances are varied so
rapidly that for any time-space point, {t, x}, there is a vicinity where we
can neglect the time and space variations of the coefficients in the
equations in variations. Then we can considerably simplify the problem,
reducing it to the local analysis of linear partial differential equations
with constant coefficients. This is the idea of the method of 'frozen
coefficients', which we employ further to analyse the Hadamard stability
of constitutive equations. Of course, by this very simplified method we
can only locally investigate extremely rapid instabilities but these types
of instability are deeply related to some fundamental properties of
partial differential equations and, as shown below, to thermodynamics.

62

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

According to the method, the disturbances are searched in the form of


periodic elementary waves: 41 97

{be, c5v, c5p} ~ 8{ C, V, p} exp[i(k x - Wt)/8 2]

(3.18)

Here 8 is a small amplitude parameter that is considered further in the


limit 8-+0; C, V and p are the complex-valued amplitudes dependent on
'slow' variables 8X and 8t; and the exponential factor in the square bracket
represents a rapidly varying phase of oscillations. According to the
method, the parameters of oscillations, wave vector k and frequency w
are assumed to be constant. A more general approach related to the
nonlinear geometro-optical approximation 144 was also considered In
Ref. 41.
The wave numbers k i will be treated as real and the frequency of
oscillations w will be searched for. Then, due to eqn (3.18), the basic
solution is stable if 1m w:::;O. It is particularly true if the frequency w is
real. If 1m w>O, the very rapid, 'explosive' instability will happen in the
limit 8-+0 under study.
Substituting eqns (3.17) and (3.18) into eqns (3.l4H3.15) and taking
into account the terms of lowest order of 8, 0(8 -1) yields:

+ 1/2(1 + ~)(iiAmkm + CirnVrnkj)


-1/2(1-~)(cirnVrnkj + kjvrnc rn ) =0

nCij

(3.19)

(n=w-vjkd
where n is the frequency of oscillations with Doppler's shift taken into
account. Equations (3.19) display the linear set of homogeneous algebraic
equations relative to the amplitudes of oscillations marked by overbars.
The same set holds for the complex conjugated amplitudes {c*,v*,p*}.
The remarkable feature of this set is that the dissipative term f(c, ~) is not
represented in these equations, because, due to the expression of disturbances in the form of eqn (3.18), the out-of-derivative terms in the
equations in variations are of the order of 0(8). This corresponds exactly
to the method of characteristics.
Consider first the simple case when O'e(c) is given by the neo-Hookean
relation, 0' e = (fl/ ~)c, as in the constitutive equations of Gordon-Schowalter type (eqn (2.10)). Then the term RjjnlCnlkj in the first equation (3.19)
is simplified to (flg)cijk j Excluding the quantities Cjj from the first and
third equations (3.19) and using the incompressibility condition, for the

Simple Constitutive Equations for Viscoelastic Liquids

63

velocity disturbances, vjk j = 0, results after simple manipulations in the


'dispersion' relation:
(3.20)
where
(3.21)
and k and Ivi are the moduli of wave vector and disturbance of velocity
amplitude, respectively. Because the Doppler shift has no effect on 1m w,
the basic solution is stable if and only if the right-hand side of eqn (3.20) is
positive semi-definite. But the right-hand side of eqn (3.20) is represented
as an algebraic sum of quadratic and Hermitian forms of independent
variables Kj and 1'fj. When ~ = 1, eqn (3.20) shows that the problem is
Hadamard stable if the configuration tensor c is positive (non-negative)
definite. However, when 0 < ~ < 1, it could be easily proved that the model
is Hadamard unstable regardless of the sign ~ if the maximum principal
value of tensor c is more than (l_I~I-l). Though for the models,74-76
shear flows stable in Hadamard sense, these models are unstable in
simple elongation. Thus, except for equations with lower- and upperconvected derivatives, the viscoelastic constitutive models (2.10) of the
Gordon-Schowalter (or Johnson-Segalman) type are generally unstable
in the Hadamard sense. It should be also mentioned that the complete
analysis of this instability by using the method of characteristics was
given first by Dupret and Marchal. 138
3.3.2 Equivalence between Hadamard and Thermodynamic Stabilities for
Maxwell-like Constitutive Equations with Convected Derivatives

Having established that even in the neo-Hookean case, the set of


equations (3.14H3.16) is generally Hadamard unstable when ~ i= 1, we
cannot expect that the nonlinearities for general dependence O"e(c) in eqn
(3.1) will stabilize the situation. Thus to analyse the effect of nonlinearity
in the dependence O"e(c) we consider further, only the case ~ = 1 in eqns
(3.19). In this case, in much the same way as it was done before, we can
reduce eqns (3.19) to the dispersion relation:
(3.22)
where
(3.23)

64

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In eqn (3.22), the variables l1i and Ki, subordinated to the incompressibility constraint l1iKi =0, are the same as defined in eqn (3.21), and the
rank-four tensor R is defined in eqn (3.15), but now with ~ = 1.
As seen from the subsequent analysis, a generalized Hermitian form
in eqn (3.22) is real valued. Then eqn (3.22) shows that the problem
is Hadamard stable if and only if the symmetric rank-four tensor B,
defined by eqn (3.23), is positive semi-definite. This is the general
evolution condition and all further analysis relates to expressing this
condition in terms of constraints imposed on possible forms of free
energy f.
Let us now demonstrate that the evolution condition obtained is the
same as the condition of thermodynamic stability for the viscoelastic
liquids under study. First consider a 'small' vicinity characterized by the
small tensor bc for a fixed stressed state with the tensor c. Here be is
represented by:
(3.24)
and fJ is the increment in the Hencky deformation tensor which is a 'true'
thermodynamic variable in Eulerian description. 52 If the constraint
detc= 1 is imposed, the additional constraint trfJ=O results. The viscoelastic liquid is said to be thermodynamically stable in the vicinity of the
stressed state c if the difference between elastic potentials in c + bc and c
'states' and the work P made by stress is positive. This definition reduces
to the following inequality:
A=Po[f(c+bc)- f(c)]-P>O

(3.25)

It is easy to show that P = 2 tr(1 e fJ) and that


Po[f(c+bc)- f(c)] =P+B ijmn !3ij!3mn

where the tensor B is defined by eqns (3.22) and (3.23). Then the condition
(3.25) of the thermodynamic stability is reduced to
(3.26)
Inequation (3.26) can be satisfied for arbitrary fJ only if the tensor B is
positive definite. This proves the equivalence between thermodynamic
and Hadamard stabilities for the case considered.
Godunov 145 ,146 was seemingly the first who completely studied the
constraints imposed by thermodynamics on the well-posedness of equations of gas dynamics. He also found necessary evolution conditions for
the set of equations valid for isotropic elastic solids, and formulated these

65

Simple Constitutive Equations for Viscoelastic Liquids

as some constraints imposed on the free energy.10 Previously Truesdell 147 obtained the same constraints which in an incompressible case are
employed as in eqn (1.28) in Chapter 1. For incompressible elastic solids,
the constraints imposed by the condition (3.25) of thermodynamic stability are studied in Ref. 8. It should also be noted that the results of the
study of Hadamard stabilities for viscoelastic liquids considered here, are
comparable with the results of studies of the stability of static and
dynamic deformations for elastic solids (see, e.g. Refs. 148 and 149).
3.3.3 Constraints Imposed on Possible Forms of Elastic Potential by
Hadamard (Thermodynamic) Stability

Now we find the expressions for the components of tensor B. Let us


consider first the case when detc#constant. Straightforward but tedious
calculations due to eqn (3.23) with allowance for eqns (3.15) and (3.1) result
in some awkward expressions for the components of B. Due to the invariant
character of the basic form (3.22), we can always use the principal axes of
tensor c where these components are simplified to the expression:
(3.27)

Here
(3.28)

and
1
-4 Lij=f2 CjCj+ f3 13+ fll CjCj+ f22 CjCj(Il-CJ(Il-Cj)
Po

+f331~+ f12CjCj(211

-Cj-Cj)+ f1313(Cj+Cj)

(3.29)

+ f23 13[Il(Cj+Cj)-cf- cf]

In eqns (3.28) and (3.29), Cj represents the principal values of tensor c and
fj = 0f/o1j, fij = a2f/iJIjolj.
Substituting expression (3.27) for tensor B into the basic form (3.22)
represents it as the sum of two independently vanishing Hermitian forms:
Q2/k 2 =

L Gj.il~jllj+~jlld2+ LIXijXjxj
j,j

(3.30)

i<j

Here
i=j
i#j

(no summation!)

(3.31)

66

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Equations (3.30) and (3.31) show that the basic form is always real-valued.
In terms of stability, the independence of two Hermitian forms (eqn (3.30))

means that they should be positive semi-definite separately. Also, due to


the incompressibility condition for disturbances, LiKi'1i=LiXi=O, the
variables Xi in the second Hermitian form are linearly dependent. Taking
into account these considerations and using common algebraic results/ 50
we can represent the criteria of positive semi-definiteness for the basis
form (3.30) as follows:
Ci>O;

Gij~O

(3.32a)

(i))

Kij=2(Gii+Gij)+Lii+Ljj-2Lij~0

(ii=j)

KijKjk -(2G jj + L jj + Lik - Lij - Ljk )2 ~ 0 (i i= j i= k)

(3.32b)
(3.32c)

Here Gij and Lij are represented by eqns (3.28) and (3.29), respectively.
Inequations (3.32) are th~ necessary and sufficient conditions for
Hadamard stability. It means that violation any of the inequalities results
in instability.
If the free energy I is such that ineqns (3.32) are satisfied in the total
physical region of Ci definition: 0 < Ci < 00, independently of the flow, we
call the model of elastic liquid globally stable in the Hadamard sense, and
stable in a region (or absolutely unstable) otherwise.
As a useful example, let us consider a particular case when the free
energy 1= 1(1 1 ), i.e. it depends only on the first invariant of the
configuration tensor c, as is proposed in molecular models. Then the
conditions of evolutionarity (3.32) are reduced to the following:
Ci~O;

1'~O;

(Ci +Cj)1' +(Ci -cY J" ~O (i i= j)

(3.33)

Here l' and J" are two derivatives with respect to 11 , Also, easy
analysis,41 based on ineqns (3.32), shows that the Mooney potential (eqn
(1.28)) is globally stable in the Hadamard sense if (and only if) Ci > 0, f.1 > 0
and O~a~ l.
Let us now briefly discuss the case detc = 1, considered in Chapter 1. Here
the evolution behaviour is just the same as for the elastic incompressible
solids and the same analysis results in the expression (3.27) for the tensor B
but the quantities Gij and Lij are now changed for the following:
GQ= W1(Ci+CJ+ W 2 (Ci- 1 +Cj-l)

(3.34)

LQ=4Wl1CiCj -4W12(Ci-1Cj + CiCj-l)+4W22Ci-1Cj-l

(3.35)

Simple Constitutive Equations for Viscoelastic Liquids

67

Here Wj =oVJ1oI j and Wij=o2W/ol j ol j The same criteria (eqn (3.32)) of


evolutionarity are also valid for the variables (3.34) and (3.35), and the
equivalence between the thermodynamic and Hadamard stabilities has
been established in Ref. 8.
Since in the case under study the matrix L~ is represented by a bilinear
form with respect to Wjj , some easier sufficient conditions of evolutionarity have also been established. s
The results on evolutionarity can be easily generalized, to some extent,
for the multimodal case with independent modes where

f(T,

Cl, C2,''''

cn )= Ifd T, Ck)
k

with equations of type (3.6) holding for each sub-tensor Ck'


The same analysis of Hadamard stability results in conditions of
positive semi-definiteness for the matrix B=I:.Bk where the submatrices
Bk have been defined earlier for the single mode. But now, using the same
stability conditions as in eqn (3.32) for each mode provides us only with
sufficient conditions for Hadamard or thermodynamic stability for the
whole model.
3.3.4 Hadamard Stability of Single Integral Constitutive Equations
The Hadamard stability of single integral constitutive equation (2.15) and
(A.45) has also been generally analysed by the method of frozen coefficients. 151 The complete analysis of the global stability was, however,
achieved only for the so-called time-strain separable case, for which the
integrand in eqn (2.15) is represented in the form:
1= g(t- r)E(c);

g(t - r) = dm(t - r)/dr

(3.36)
Here C = C- l is the (positive defined) Finger strain tensor and m(t) is the
relaxation function with properties (1.90), which guarantee the existence
of instantaneous elasticity. It should be noted that the time-strain
separability was found experimentally, and is now widely accepted 6 in
viscoelastic modelling.
The main purpose of the study described in Ref. 151 was to establish
the necessary and sufficient conditions in the general case when the
material functions <Pl and <P2 are not potential, i.e. do not obey the
conditions (2.16). In much the same way as was done for general
Maxwell-like constitutive equations, it was found that the conditions of

68

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

global Hadamard stability for the case under study, are expressed in the
form:
(3.22a)
with the basic form reminiscent of that expressed in eqn (3.22). Here the
symmetric rank-four tensor Bijmn is represented in the form of eqn (3.23),
where the rank-four tensor R is changed for R given by:
(3.15a)
and tensor E being defined in eqn (3.36). Because met) is a posItIve
function, the condition of Hadamard stability (3.32a) is now reduced to
positive semi-definiteness for the tensor B. These in turn, are reduced to
the same criteria stability (3.32) in the principal axes of tensor c, with the
functions Gij and Lij changed for the following:
(3.28a)
iij

= 4CPll CiCj + 2CP12 CiCj (21 1 -

2CP22 Ci- 1 Cj-l

Ci -

Cj) - 2cp21 (CiCj-l + Ci- 1 Cj)

[I 2(Ci + Cj) - 2] - 2CP2(Ci + Cj)(1 1 -

Ci - Cj)

(3.29a)

Here CPmn == ocpm/o1n (m, n = 1, 2).


In the potential case, CP12=CP21 (CPl and CP2 are proportional to W 1 and
W 2 , respectively), eqns (3.28a) and (3.29a) are reduced to eqns (3.34) and
(3.35), respectively, which means that in this case, the same equivalence
between thermodynamic and Hadamard stabilities holds true as has been
established for Maxwell-like constitutive equations of a differential type.
No general results have yet been established for the non-potential
constitutive equations where CP12 =1= CP21. The results of analyses of
Hadamard stability for some particular cases of eqns (2.15H3.36) are
presented in Section 3.4.
3.3.5 Positive Definiteness of the Configuration Tensor:
'Dissipative' Instability

Unlike single integral constitutive equations, where tensor C is always


positive, the evolution behaviour of general Maxwell-like constitutive
equations (3.1) and (3.6) essentially depends on the positive definiteness of
configuration tensor c. As follows from the results of Section 3.3.3, if in
any point of the medium a principal value Ci is negative, the Hadamard
instability will immediately happen. Moreover, in this case, we cannot
also satisfy the restrictions imposed by the Second Law. Thus along with
non-evolutionarity, heavy 'dissipative' instability can also happen. Some

Simple Constitutive Equationsfor Viscoelastic Liquids

69

sufficient conditions for the configuration tensor to be positive definite


have been found by different methods in Refs 41 and 152. Nevertheless,
the same results were obtained in both references, and can be summarized
as follows:
(a) Theorem 141
For any given, piecewise smooth strain history, the configuration tensor c
which satisfies eqn (3.6) and initial condition at the rest: cl t = 0 = 15, is
positive definite.
Here the piecewise smooth strain history is defined as almost smooth,
i.e. except for discrete points and surfaces where it suffers by some limited
discontinuities. At these singular points and surfaces, the velocity gradient tensor has the Dirac o-components. Theorem I was proved in Ref. 41
by studying the properties of the solution of eqn (3.6) represented in the
principal axes of tensor c, whereas in Ref. 152 it was proved by analysing
the general representation of dissipative term ljJ(c) in eqn (3.6). In spite of
the limited character of Theorem I, which will be discussed in Section 3.5,
it can guarantee the Hadamard stability of the viscoelastic constitutive
equations in many flows.
Neither positive definiteness of tensor c nor satisfying the conditions of
Hadamard stability (3.31) or (3.36) can guarantee, however, the absence of
another type of instability related to poor modelling of dissipative term
ljJ(c) in general Maxwell-like equation (3.6); so-called dissipative instability. The familiar example of instability of this type is the simple extension
for the upper-convected Maxwell model where, without any reason, the
inertialess solution of the problem goes to the infinity when the product
of elongation rate by relaxation time is more than 1. This artefact has
been removed in molecular approaches by introducing the FENE potential (for details, see Refs 5 and 6). But it is still unclear what constraint
should be imposed on the dissipation term ljJ(c) in eqn (3.6) in order to
guarantee the dissipative stability. To some extent, the following theorem
gives sufficient conditions for the dissipative stability and boundedness of
the tensor c.
We will call a second-rank tensor h 'limited' if its norm, Ihl =(trh2)1/2 is
bounded with a positive constant, and call a flow of liquid 'regular' if the
velocity gradient tensor is limited, whether the flow is stationary or not.
(b) Theorem JI41
Consider the set of Maxwell-like constitutive equations (3.1) and (3.6)
with positive dissipation D=D(T,1 1,1 2,!3) defined in eqns (3.4) and

70

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

(3.6). Let the free energy f be a not-decreasing smooth function of


three invariants I k If, for any positive number A, the asymptotic
inequality
(3.37)
holds, then in any regular flow, the configuration tensor c and stress
tensor (1e are limited.
The proof of this theorem uses the fact that due to Theorem I, tensor c
is positive definite for any regular flow. Theorem II holds whether the
elastic liquids are considered as compressible or not and independently of
the assumptions about det c. This theorem can be useful for modelling the
dissipative term ljI(c).
At present, almost nothing is known about dissipative instability in
single integral constitutive equations.
3.4 EXAMPLES
This section mainly considers some examples of Maxwell-like constitutive
equations for the incompressible viscoelastic liquids that have been
discussed in Section 2.1. Here three topics are under discussion: (i)
representing the equations in the 'canonic' form of eqns (3.1) and (3.6),
including the expression for the free energy; (ii) the conditions of
Hadamard stability; and (iii) the relation of the expression for dissipation
to the conditions of Theorem II. The predictability of the equations is
discussed, for example, in Ref. 6. At the end of this section, the results of
analyses of Hadamard stability are discussed for some particular cases of
time-strain separable single integral constitutive equations (2.15H3.36).
3.4.1 Upper-Convected Maxwell Model
The upper-convected Maxwell model is presented in eqns (3.1) and (3.6)
with
ljI(c) = (c- i)/8 = 0

The elastic potential, W

(3.38a)

Po f, and dissipation, D, are:

W = /-l/2(Il - 3),

D=W(Il -3)/(28)

(3.38b)

Equations (3.38b) show that the elastic potential is neo-Hookean. The


expression for the dissipation is discussed in Refs 84 and 138-142.

Simple Constitutive Equations for Viscoelastic Liquids

71

As known (and demonstrated once again in Section 3.3), this model is


globally Hadamard stable. For this model,

I(lel = Gltre 2 = Gm - 212)1/2


1

(3.39)

Thus eqns (3.38b) and (3.39) show that the sufficient condition of
Theorem II, ineqn (3.37), is violated. This is the reason for the unlimited
behaviour of this model in simple elongation and other, more complicated flows, resulting in the appearance of dissipative instability.
3.4.2 Dumbbell-spring Model with Finite Extensibility
The easy derivation shown in Ref. 6 (Section 8.5.3), results in the canonic
form of eqns (3.1) and (3.6) where:
(Ie

= JlKe,

t{!(e) = (Ke - (j)/8

K(11) =(1e - 3)/(1e - 11)

(3.40a)
(3.41)

Here Jl=nkT is the elastic modulus, e=<RR)/<R6), 1e=R~=constant


and R is an end-to-end vector with mean-square equilibrium length Ro
and finite length Re. In order to satisfy the transition (2.12) to equilibrium
conditions, the numerator, 1c - 3, is included in the expression for
function K in eqn (3.41), which was absent in original papers (see, e.g.
eqns (8-61) and (8-62) in Ref. 6). The elastic potential Wand dissipation
D for this model are expressed as follows:
W = (Jl/2)(1 e - 3) In K,

As followed from ineqns (3.32), this model is globally Hadamard stable


and has limited variables in any regular flow.
3.4.3 Phan-Thien/Tanner Upper-convected Models
These models take the canonic form of eqns (3.1) and (3.6) where
(Ie

= Jle,

t{!(e)=(e-(j)g(11-3)/8

(3.42a)

and scalar function g is either linearly or exponentially increasing when


11 - 3 increasing. The elastic potential and dissipation are represented as
follows:
(3.42b)
and show that this model is globally Hadamard and dissipative stable.

72

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

3.4.4 Larson Model


The Larson model has been discussed in Section 2.1. Its canonic representation which summarizes eqns (2.14) is:
(Ie

= JiC/B(I d,

W = (3p/2rx) In B(Il),

t/I(c) =B(I d(c-15)/8


D = (p/28)(I 1 - 3)

B(I 1)= 1 + (rx/3)(I 1- 3)

(3.43a)
(3.43b)
(3.44)

Here )1. is the elastic modulus, 8 is the relaxation time and rx is a numerical
fitting parameter (O:::;;rx:::;;l). Now, by using the criteria ofevolutionarity
(eqn (3.33)) valid for the case, we find that the model is globally
Hadamard stable. Also,

I(lel =w(Ii -Iz)l/z/B(Id


and the condition (3.37) of Theorem II is satisfied. Thus the model has the
limited variables c and (Ie for any regular flow.
3.4.5 Giesekus Model
The Giesekus model has been formulated in Ref. 6 in the canonic form of
eqns (3.1) and (3.6) where
(le=)1.c,

t/I(c) = [15 + P(c-15)] (c-15)/8

(3.45a)

where O:::;;P:::;; 1. The elastic potential is neo-Hookean, i.e. W = ()1./2XI 1 - 3)


and the dissipation has the form:
(3.45b)
In the principal axes of tensor c, the dissipation can also be represented as
follows:
D = ()1./28) L [1 + P(Ck -1)](Ck -1)

(3.45c)

Thus this model is globally Hadamard and dissipative stable.


It should be mentioned that at this time, there is no proof of positive
definiteness for dissipation neither for the general Maxwell-like model
(3.1) and (3.6) nor for the specific models under discussion.
Now, following Ref. 151, we briefly consider the Hadamard stability of
some particular versions of single integral, time-strain separable constitutive equations (2.15H3.36) specified by the choice of material functions
<fJk(I 1 ,1 2 ) in eqn (3.36).

Simple Constitutive Equations for Viscoelastic Liquids

73

3.4.6 Oldroyd-Lodge ModeI 69 72


Here,
C{J1 = 1,

(3.46)

This model is globally Hadamard stable.


3.4.7 Wagner Model 191
Here,
C{J1 =h(I 1,1 2 )= j exp( -n1JI=3)+(1- f) exp( -n2JI=3)
C{J2 =0; I = PI 1 +(1- P)I 2
(3.47)

and j, n1, n2, and P are positive numerical fitting parameters.


The studies of stabilities of simple shear and simple elongation 151
showed that in both regimes, the criteria of Hadamard stability (eqn
(3.32 are violated, when deformations are intense enough. This happens
even in the potential case P=O. Thus this model is globally Hadamard
unstable.
3.4.8 Wagner Model 1192
Here
and 0( and P are positive numerical fitting parameters. We exclude
from the consideration the case 0(=0, when the Wagner model II
coincides with Kaye-BKZ model with Money-Rivlin potential, which is
considered later. Then the study of simple shear 151 exposed the
Hadamard instability under intense shearing, if P~+ In the potential case
P=t, the Wagner model II is Hadamard stable in simple shear. Thus,
with non-characteristic exceptions, the Wagner model II is globally
Hadamard unstable.
3.4.9 Kaye-BKZ Model with Mooney-Rivlin potential (eqn (1.46
Here,
C{J1=1-0(,

C{J2=0(,

(3.49)

A brief analysis of criteria (3.32) of Hadamard stability shows that this


model is globally Hadamard stable if and only if, 0,,;; 0(";; 1. The same
result was obtained in Ref. 41 for Maxwell-like constitutive equations
(3.1) and (3.6) with Mooney-Rivlin potential.

74

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

3.5 DISCUSSION
In this section, we summarize and discuss the effects found in the previous
sections of the chapter. These effects are mainly related to a comparison
between the fundamental properties of the model developed in Chapter 1
and for other differential constitutive equations discussed in this and
previous chapters.
3.5.1 The Importance of trep for the Formulation of Maxwell-like
Constitutive Equations
As shown above, the general local formalism of irreversible thermodynamics results in simpler derivation of Maxwell-like constitutive
equations and their generalizations with the same results as obtained by
Poisson-bracket formalism, regardless of whether the constraint tre p = is
employed or not. By using the same method we could also easily derive
the kinetic equation for the evolution of tensor c with mixed upper- and
lower-convected derivatives bearing in mind, however, that these models
have no limit transition to the model of perfect elastic solids and that
these are Hadamard unstable. Thus below, we will discuss mostly the
nonlinear Maxwell-like viscoelastic constitutive equations (3.1) and (3.6).
The derivation shown in this chapter for both compressible and
incompressible cases, demonstrates a principal difference between the A
models, which ignore the constraint trep = 0, and the B models, which
employ it. While in the A models, the density is not described by the state
variable, configuration tensor c, which the free energy depends on, in the
B models, the density is described by the tensor c, exactly as in the case of
elastic solids. Moreover, in the A models, the configuration tensor c is a
typical non-measurable internal parameter, whereas in the B models, it
has the sense of recoverable strain and can, in principle, be measured
independently.
In the incompressible case, the condition tre p = results in the constraint 13 = detc = 1 which the A models ignore and the B models use
extensively. In the A models, ignoring this constraint results in the situation
when in general, the thermodynamic stress (1e is defined by the same
equation (3.1) as in the compressible case, and differs from the actual
stress (1 by non-equilibrium isotropic pressure. But this contradicts the
very definition of the Maxwell models. In contrast, the actual stress
tensor (1 in incompressible B models, coincides completely with the
thermodynamic stress (1e which is defined, as in equilibrium, by the
familiar Finger formula (1.24) with the accuracy of isotropic equilibrium

Simple Constitutive Equations for Viscoelastic Liquids

75

pressure. For the B models, where the constraint 13 = 1 holds, it is easy to


prove, as in the case of the incompressible elastic solids, that 1 1 > 3 (and
also 12 > 3) if tensor c is positive definite, but we could not prove this for
the A models. Yet in the first four examples of Section 3.4 for the A
models, the dissipation is proportional to the term (11 - 3), not mentioning the Giesekus model where the dissipation is defined by a more
complicated expression.
Thus the very difference between the A and B models lies in the values
of 13 Many of the A models were derived by using molecular approaches
discussed in Section 2.3 of this book. In these derivations, the magnitude
/2 stands for Jacobian of local transformation of actual to equilibrium
(at the rest state) configuration spaces. In the case of incompressible
rubber elasticity, this Jacobian is equal to unity due to the affinity
assumption. When starting to consider the behaviour of a dilute polymer
solution within, say, the dumb-bell approach, where the kinetic equation
for the distribution function involves elastic, viscous and Brownian forces
acting on a single dumb-bell, there is seemingly no reason to use any
constraint imposed on the Jacobian. However, when dealing with concentrated polymer solutions and melts, one has also to take into account
such collective effects of macromolecule interactions as side repulsion of
the chains, which have not been considered in the simplified molecular
approaches like the 'strand' approach.6 Perhaps in this case, and especially for polymer melts, the constraint 13 = 1 is a good approximation to
take into account the side repulsions as in the case of rubber elasticity.
When the tensor c has the sense of Finger recoverable strain, there is
no doubt that the condition of incompressibility 13 = 1 is valid. So it puts
the rheological modelling of elastic liquids on the solid basis of continuum mechanics and thermodynamics where one can use in the elastic limit
the stress-strain dependences studied in the rubber elasticity. For this
case of modelling elastic liquids, 12 = trc- 1 and 13 = 1, so we can involve
both the invariants 11 and 12 in the modelling. In contrast, in all the A
models only the invariant 11 has been involved in the description of
stresses, seemingly because the straight molecular sense of other invariants is vague (as in the physics of rubber elasticity). This situation limits
our abilities to model the complicated properties of elastic liquids.

3.5.2 Hadamard and Dissipative Instabilities


Now let us discuss the problems associated with the possible unstable
behaviour of viscoelastic constitutive equations. As mentioned, there are
two obvious possible origins of instabilities in viscoelastic constitutive

76

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

equation of differential type: one is related to the modelling of elastic


terms in eqns (3.1) and (3.6) which could result in Hadamard instability
(or non-evolutionarity, or ill-posedness), and another, dissipative instability, is related to the modelling of dissipative terms in eqn (3.6).
There are contrary opinions in the literature about the physical
importance of the non-evolutionary behaviour of rheological equations.
Joseph and co-workers 153 considered the loss of evolution in viscoelastic
constitutive equations as physically acceptable and speculated that this
might be related to the melt flow instabilities observed in flows of
polymeric liquids. By using slightly stabilized constitutive equations,
Malkus and co-workers (see e.g. Refs 154 and 155) developed a description of spurt flow as a trigger mechanism for the melt flow instabilities. In
contrast, Dupre and Marchal 138 considered the non-evolutionary behaviour in the more traditional way usually employed in physics and
mechanics, as absolutely unacceptable.
The problems of Hadamard (and other) stabilities can be and have
been investigated without any relations to thermodynamics. But the pure
mathematical studies, however sophisticated they are, cannot provide us
with a physical sense of phenomena. Thermodynamics seems to be a
promising tool for that. A broad class of viscoelastic constitutive equations with instantaneous elasticity, that are related to thermodynamics,
has been established. Fortunately, almost only this class has been widely
tested and used for modelling flow phenomena in viscoelastic liquids. It
was proved that within this class, the equivalence exists between the
Hadamard and thermodynamic stabilities. It means that at least for this
class, ill-posedness means thermodynamic instability without any physical reason. Therefore the Hadamard unstable constitutive equations from
this class should be rejected, simply as poorly formulated. No general
results of global Hadamard stability for the non-potential viscoelastic
constitutive equations have been obtained. But the examples of Wagner
models, demonstrated in Section 3.4, showed that the non-potentiality
leads to ill-posedness (in the global sense). Moreover, as shown in the
example of the Wagner model I, the potentiality itself does not guarantee
the well-posedness; only the conditions of thermodynamic stability do
that.
Yet it has long been known, starting from the early results of
Hadamard instability 145,146 (see also Ref. 97) that adding a viscous term
with an arbitrarily low viscosity, stabilizes a Hadamard unstable system.
But when the viscosity is very low, the smoothed unstable terms with
unclear physics dominate and result in some self-oscillations, which is the

Simple Constitutive Equations for Viscoelastic Liquids

77

physical reason why only the upper-convected derivative in the kinetic


equation (3.6) for tensor c is considered in Refs 8 and 9, as well a.s
throughout the present book. In this case, as shown in Ref. 41 and
Section 3.3 of this book, the conditions of Hadamard stability are exactly
the same as those for the thermodynamic stability of elastic solids, and
result in some constraints imposed on possible forms of free energy.
As mentioned previously, for the constitutive equations under study,
dissipative instability could also occur, as in the case of the upperconvected Maxwell model. Sufficient conditions for a given flow history
were found and formulated as Theorem II in Section 3.3.4 to avoid the
unlimited behaviour of rheological variables associated with the modelling of dissipative terms in constitutive equations. For all the examples
considered in Section 3.4, these conditions were satisfied. Yet for all of
them, there is no proof that the dissipation is positive definite in all the
admissible regions of variation of the configuration tensor.
Both theorems in Section 3.3.4 have, however, a limited character
because they are based on the assumptions made about the flow history.
But in the real situation, the flow field has to be found as a solution of the
initial boundary problem for the total set of equations, including the
momentum balance. Thus in a computational process we usually deal not
with a given flow history but rather with mixed, flow-stress history where
the theorems do not work. One example of unlimited behaviour for the
Giesekus and simple Leonov model is demonstrated in Chapter 5 for the
simple shear flow when the shear stress is given.
The formal method of 'frozen coefficients' is not applicable to the
situations when a system is described by a set of differential equations
that are nonlinear in derivatives for the basic variables c, v and p. The
White-Metzner model 78 described by eqns (2.1) and (2.5), is an example of
such a system. The Hadamard instability for this model was proved in
Ref. 138 by the method of characteristics. Also, the Hadamard stability of
viscoelastic constitutive equations of general integral type can be studied
by using the rate equations (see, e.g. Sections 15.3 and 15.4 of Ref. 97).
Many very general results can be obtained by using this approach. But
the constructive results for global Hadamard stability, with the stability
criteria expressed through material functions but not functionals, can
seemingly be achieved only for time-strain separable single integral
constitutive equations. In this case, the method of frozen coefficients
trivializes the problem to infinitesimal and local stability analysis, and the
same results can be obtained by using more general methods.

CHAPTER 4

Experimental Methods Tn the Rheology of


Viscoelastic Liquids

4.1 INTRODUCTION
The rheological properties of viscoelastic polymeric liquids, which in the
case of polymeric melts and solutions, are highly viscous, are mainly
tested in very simple flows called viscometric flows and elongational flows.
With experimental data obtained by studying these flows, one can easily
and reliably evaluate the descriptive ability of various constitutive equations, define material constants and determine the differences between
materials in technological processes, etc.
The most common of the viscometric flows is simple shear with
capillary flow, with flows between cone and plate, rotating disks or
coaxial cylinders, being particular examples. The most familiar of the
elongation flows is uniaxial elongation. These two types of rheological
test are widely used for liquid polymers and are carried out under various
conditions, duplicating each other only in the linear region of deformations.
Historically, viscometric methods were developed for studying the
viscous properties of liquids, and the elongational methods, for testing
solids. However, when dealing with polymer liquids, often one has to take
into account their elastic (solid-like) properties in the shear flows and also
their liquid-like features in steady elongational flows.

79

Rheology of Viscoelastic Liquids

First, we shall briefly discuss the viscometric methods that have


received a lot oftreament in the literature 7 ,lOl,156,157 and then give more
attention to the elongation tests developed in the last 20 years.
4.2 SIMPLE SHEAR TESTS
4.2.1 General Relations

In simple shear, the velocity has the following components:


(4.1)

V2=V3=0

where y is the shear rate, the axis Xl is directed along the flow, the axis
X2 being orthogonal to Xl is disposed in the shear plane, and the axis X3
is orthogonal to the shear plane. The strain rate and vorticity tensors are
represented by
e=

0
(

-1o 0)

1 0

0
0

(4.2)

and the stress tensor is: 10l


(4.3)

(J'22

o
Usually measured quantities are:
(J'2=(J'22-(J'33

(4.4)

where (J'12 is the shear stress, and (J'l and (J'2 are thejirst and second normal
stress differences, respectively. When elastic liquids are considered as
incompressible, both the normal stress differences bear 'pure' rheological
data since in this case, the pressure p is excluded from the consideration.
For the Newtonian liquids, (J'1=(J'2=0, and the non-zero normal stress
differences reflect elastic properties of these.
When shear rate and stresses are independent of spatial coordinates,
the shear is called homogeneous, and inhomogeneous otherwise. Only the
homogeneous shear rate and the respective instruments for its study are
considered in the following parts of Section 4.2.

80

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

4.2.2 Parallel Plate Device


The deformation of simple shear can be established in the simple way when
a liquid flows between two parallel plates (Fig. 4.1). One of the plates is fixed
and another moves with a velocity v which generally depends on time t. The
shear rate is defined here as follows: y= viH, where H is the gap width
between two plates. Devices of this type are called plastometers.
Experiments with plastometers are practical only with small strains, i.e.
in the linear region of deformation where normal stresses are insignificant.
This is due to increasing edge effects at higher strains. In these experiments,
there is sometimes an opportunity to release the mobile plate in the course
of the experiments and measure the recoverable strains. Usually the upper
plate moves under the action of a free load connected with the plate by a
thread via a roller (dashed lines in Fig. 4.1). The load has to be high as
compared to the frictional force created by the thread on the roller. Also, the
vertical motion of the upper plate under its weight should be negligible. In
the course of experiments, the device is placed in an air thermostat. The
slow travel of the mobile plate along the Xl axis is recorded, for instance, by
means of a microscope which follows the displacement of a needle
connected rigidly with the plate, along a ruler.
This apparatus can be also used for studying the strain recovery
process. In this experiment, the load is suddenly removed and the upper
plate moves under the action of elastic energy stored during the predeformation period in a direction opposite to that before retardation. It
should be noted that if normal stresses are important, the mobile plate
will not only have a reverse tangential, but also an upward motion, along
the X2 direction.
Plastometers are convenient for high viscosity polymer melts to
measure their maximum Newtonian viscosity 1J and the equilibrium
high-elastic modulus Ge (see eqns (1.90) and (1.91)).

c::r=t:::::::~;::=r=::J----

---(7-.... ,

!----'.<.,.<----"----_,z,

\,1

I
I

o
I

Fig. 4.1. Simple shear: parallel plate device.

Rheology of Viscoelastic Liquids

81

To study the nonlinear effects in steady and transition flows of polymer


liquids one could consider the Couette viscometer with flow between
rotating co-axial cylinders as sketched in Fig. 4.2. When the gap between
the cylinders, H, is considerably less than the radius of the internal
cylinder, R(R ~ H), the flow is very similar to that between two plates.
The case of Couette flow with an arbitrary gap was considered in Ref.
156. Unfortunately, the Couette device is inappropriate for highly elastic
liquids, because these flow out in the axial direction from the gap between
the cylinders due to the familiar Weissenberg effect.
4.2.3 Cone-Plate Rheometer: Basic Relations
Cone-plate rheometers are commonly used in the rheometry of highly
viscous polymer melts and concentrated polymer solutions. Figure 4.3
shows a scheme of the cone-plate rheometer. The liquid (2) under study is
deformed between the cone (3) and the plate (1). This instrument is very
popular in polymer rheology because it can provide a researcher with
easily interpreted data of the shear and normal stresses depending on
shear rate in steady state, transitional and oscillatory flows. Another
advantageous feature of this apparatus is the simplicity of charging a
highly viscous fluid into the working unit: the cone is raised, the polymer
is put on the plate, then the cone is lowered again. Sometimes for a fast
thermos tatting, the tested polymer is extruded through an orifice drilled
I

-'

--

I
I

II

I
Fig. 4.2. Cylinder--cylinder rheometer.

Fig. 4.3. Cone-plate rheometer: 1,


rotating plate; 2, tested liquid; 3, cone;
4, shaft; 5, air bearing; 6, normal stress
transducers; 7, rod; 8, bearings; 9,
beam; 10, displacement transducer; 11,
compensating device.

82

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

along the cone's axis 158 by a syringe. This is schematically shown in


Fig. 4.4.
Consider once again Fig. 4.3. Here the cone is unmovable and the plate
is rotating in the assigned manner with an angular time-dependent
velocity Q(t). Let us introduce the spherical coordinates r, qJ, 8. The axis
8 = 0 coincides with the axis of rotation. The angle 8 varies between nl2
and nI2-a, where a is the angle gap between the cone and plate. When a
is very small (commonly, ()( = 1-30), there exists with an accuracy of 0(a 2 ),
a component of the velocity:
v =rQ n12-8
tp
nl2-a
Then the strain rate

y= 2eo",

(4.5)

is defined as follows:
y~QrIH=Qla

(4.6)

Here H = ra is the gap width between the cone and plate, and r is the
current radius. Note that since a homogeneous flow is distorted near the
cone edge, eqn (4.6) holds not for the whole range 0 ~ r ~ R of r variation
but only for r ~ R* < R, where R is the cone radius and R - R* = O(Ra).
Also, in homogeneous region, r < R*,
(J9r = (Jtpr = 0

(4.7)

and the components of stress tensor


(4.8)

do not depend on the space coordinates. These assumptions are fulfilled


for simple shear independently from a rheological model.
The inertial equations of the momentum balance (the equilibrium
equations) yield (see, e.g. Ref. 157):

Fig. 4.4.

(J tp9 = M 1(2nR;)

(4.9)

o(Jrr/or + (2(Jrr- (J99 - (Jtptp)lr = 0

(4.10)

Scheme of polymer injection into the gap between the cone and plate:
1, tested liquid; 2, cone; 3, plate; 4, syringe.

Rheology of Viscoelastic Liquids

83

Due to eqns (4.8)


2arr - a99 - aq)(p =

2a2 - a 1

(4.11)

In eqn (4.9), the quantity M is a function of time t, only. If the flow


inhomogeneity in the edge region is negligible, M is equal to the torque
affecting the disk. From the last formula in eqns (4.8) and the homogeneity of the flow, it follows that
(4.12)
Integrating eqn (4.10) over r with allowance for eqns (4.8), (4.11) and (4.12)
yields:
(4.13)

Now let us calculate the thrust, i.e. the force acting on the plane along the
rotation axis. If the calculation is confined by the contribution of the
homogeneous flow, we obtain:
F= -2n foR> a99rdr

Integrating this expression by part and further substituting the obtained


result into eqns (4.12) and (4.10), gives:
(4.14)
In deriving eqn (4.14), we used the independence of normal stress
differences from the radius. If the inhomogeneous region of flow contributes almost nothing in the total thrust, we can use in eqn (4.14) the
approximation: R* ~ R. If, additionally, arrlR ~ 0, eqn (4.14) is simplified
to:
(4.15)
which gives the opportunity to find the first normal stress difference a 1
via measured thrust F.
Hitherto, we have considered the case when the cone touches the plate.
If the thrust F is measured for a steady flow at various distances b
between the cone top and the plate, the second normal stress difference
can be also measured:159.160
.
Q [
.
2F (
az('l') = Q-ay al('l')- nR2

dlnF)]
1- dlnb

(4.16)

84

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In this case, the shear rate y is defined as follows:

y= ORj(b + Rrx)

(4.17)

4.2.4 Cone-Plate: Schemes of Instruments and Methods


of Measurement

Both the law of motion and tangential stress can be found from
measurements of a rotating 'surface', e.g. that of the plate.
In order to determine the value of tangential stress 0"12 for the given
kinematics, the torque M is measured which twists the shaft (4), fastened to
the body of the instrument, with the cone (3) at the end (Fig. 4.3). The torque
is estimated by the turn angle of the shaft by means of displacement
transducers having no contact with the body of device. To maintain the
stability of the shaft (4) the body is usually mounted in an air bearing (5).
If the plate is rotating under given torque, its turning angle is measured
as a function of time.
The tangential stress 0"12 = O"tp8 is calculated using eqn (4.9) where the
small conicity is taken into account and the edge effect is neglected, i.e. it
is assumed that R = R* and that M is the torque.
For the regimes with given kinematics or torque, the normal stresses
can be measured locally by transducers (6), say, of the diaphragm type
(Fig. 4.3), that are mounted over the radius of the plate (1) or the cone (3).
The descriptions of transducers can be found, for example, in Ref. 156.
One can, in principle, measure by using these transducers the whole
radial distribution for the stress 0"88(r, t). Also, when the central region of
the gap is filled in with various volumes of a Newtonian incompressible
fluid, with the polymer remaining outside, it is also possible to measure
the dependence O"rlr, t).lS7 If the distributions 0"88(r, t) and O"rlr, t) have
been measured, it is easy to calculate also the distribution of O"tptp(r, t) by
using eqns (4.10) and (4.13). From the values of normal components of the
stress tensor, we can find both the normal stress differences, 0" 1 and 0"2;
their independence from the radius controlling the validity of the experimental data.
The first normal stress difference, 0" b is most often determined, however, from measuring the thrust F and using the approximate formula
(4.15). To measure the thrust F, the rod (7) is made fast to the disk (Fig.
4.3) and the bearings (8) allow for restricted movements of the rod in the
axial direction. The rod pressurizes the beam (9); the force F being
measured by its deflection. The same method is also common for
measuring the second normal stress difference, 0"2, with the use of eqn

Rheology of Viscoelastic Liquids

85

(4.16) (see, e.g. Ref. 161). All these experiments are usually carried out

under isothermal conditions which are attained by putting the working


unit into an air thermostat.
When the kinematics of shear are given, there are the following regimes
of deformation: (i) the regime with the constant rotation rate (constant
shear rate); (ii) the oscillatory mode known as the dynamic regime; (iii) the
regime of relaxation where the flow stops after a shearing; and (iv) various
superpositions of oscillations and shearing.
In dynamic experiments (see, e.g. Refs 66 and 162), the shear strain is
commonly given by the relation: y = e sinwt, where e is the strain amplitude
and w is the frequency of oscillations. In any simple shear apparatus,
e ~ 2/H where 2 is the displacement of a mobile plate and H is the width of
the gap between the plates. The low-amplitude experiments, where e ~ 1, are
used most often, and deliver the information about the linear region of
viscoelastic behaviour for tested liquids. In these experiments, the shear
stress is expressed as follows: a 12 = am sin(wt + qJ) where the magnitude qJ is
the phase shift between the measured oscillations of stress and given
oscillations of strain. For a pure elastic body, qJ = 0, for the Newtonian
liquid, qJ = n/2 and for a viscoelastic liquid where the phase shift depends on
the frequency, the phase shift is limited: 0< qJ(w)<n/2. Another quantity
that is measured in the course of the experiments is the stress to strain
amplitude ratio, am/e, which has a sense of a 'modulus'. There are two
frequency-dependent moduli that are most often cited in the literature:

G'(w) = (am/e)cos qJ

(4.18)

which is called storage modulus and characterizes the elastic (equilibrium)


processes in viscoelastic liquids, and

G"(w) = (am/e)sin qJ

(4.19)

which is called loss modulus and characterizes the inelastic (dissipative)


processes in viscoelastic liquids. These moduli can also be considered as
the real and imaginary parts of the complex dynamic modulus, G*(w)=
G'(w) + iG"(w). Also often mentioned in the literature, is the complex dynamic viscosity, 1/* = G*(w)/(iw), as well as its absolute value and real and
imaginary parts. The dynamic data can also be obtained by using the
Maxwell rheometer,163,164 consisting of two rotating parallel disks with
displaced axes.
When the shear stress is given, there are two particular types of
deformation: (i) transitional shearing under constant shear stress and (ii)
the regime of retardation.

86

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Now let us present one of the possible instrument arrangements 165 ,166
operating under conditions of a given constant shear stress (Fig. 4.5(a)).
Here a liquid is sheared between the cone (2) and the plate (disk) (3). The
cone (2) rigidly connected with the pulley (4) rotates under the action of
two equal weights 5. The weights are put in motion by cutting the thread
(6). The shear is determined by the relation: y=<p/rx where rx is the angle
between the cone and the plate, and <p is the angle of the cone's turning,
which is measured by employing a potentiometer mounted on the shaft of
the cone (2). Then the strain rate is given by: y= cp/rx where the overdot
means the time derivative.

S2(i)

a)

G)
Fig. 4.5. (a) A device for simple shear at given constant shear stress: 1, tested
liquid; 2, cone; 3, plate; 4, pulley; 5, weights; 6, stopping thread. (b) A device for
measuring retardation: 1, tested liquid; 2, cone; 3, plate; 4, cone.
A good device for measuring one-dimensional retardation is sketched
in Fig. 4.5(b). The liquid (1) is sheared in two symmetric gaps between the
unmovable disk (3) whose stationary position is fixed by an electromagnet, and two similar cones (2 and 4) rotating, say, with a constant angular
speed. When reaching a certain pre-shear, the disk (3) is released
practically instantaneously. At the same time, the cones (2 and 4) are
stopped. After this, the disk (3), under action of elastic energy accumulated in the pre-shearing, starts to rotate in the direction opposite to the
initial motion of the cones. The turning angle of the disk, which can be
considerable, is recorded using a contactless method, viz., a reflected light
beam (see, e.g. Ref. 165).

Rheology oj Viscoelastic Liquids

87

The results of rheometric studies are considered as correct if they are


independent of the geometry of operating units, torsion beams and the
types of viscometric flows. There are many possible errors that may
appear when taking measurements using cone-plate instruments, which
are briefly analysed in the following section.
4.2.5 Methodical remarks
In routine rheometric experiments, we can usually hope to obtain
reproducible results until such 'critical' phenomena known for elastic
liquids, as melt flow instability, detachment of polymer from the wall and
the rupture of liquid set in, or appreciably manifest themselves (see
Chapters 11 and 12). These phenomena seem to occur first, when the
polymer begins to chip out of the gap between the cone and plate,
and are commonly related to a critical value of Deborah number,
Deer'" 1. Now let us estimate the influence of secondary effects which are
usually neglected in interpretations of polymer flows in cone-plate
devices.

(a) The edge effect


When the shear rates are high enough (De'" 10), a very non-uniform
secondary flow is initiated near the edge region. A polymer liquid is
intensively chipped out of the gap between the cone and plate, which
results in distortions on the time record of measurements. The edge effect
imposes a main limitation on the tests when operating with cone-plate
instruments. Evaluation tests in this case, e.g, with y= constant, are
carried out within the minimum time t'" () until the flow becomes nearly
stationary. When such tests do not work, they are performed using other
visco metric devices, e.g. capillary viscometer.
(b) Inertia effects
These supposedly small effects can be estimated when substituting the
inertialess expressions in the neglected inertia terms and comparing these
with preserved terms in the equations of momentum balance. For
instance, in the steady-state cone-plate flow, the inertialess approach is
valid only if
(4.20)

where p is the density, n is the disk angular velocity, R is the disk radius
and 0"1 is the first normal stress difference. Here the common inequality
10" 21 ~ 1 and the empirical relation, 0" 1'" 0" 12 have been taken into account;

88

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

(12 being the second normal difference and (112 being the shear stress.
More detailed methods for inertia estimations and corrections can be
found in Ref. 157. The inertia effects can be reduced significantly if it is
possible to decrease the values Rand n under fixed values of y.
(c) Effects of elasticity of torsion bars

Sometimes an elastic twisting of the torsion bar on which the cone is


usually mounted and which indicates the torque, can significantly influence the flow of viscoelastic liquids. This can happen in such timedependent situations as intense transitional flows, relaxation, retardation,
high frequency oscillations, etc. In this case, the supplementary characteristic time t* appears
t*=R 3 Yf/K

where Yf is the Newtonian viscosity of elastic liquid, K is the rigidity of a


torsion bar related to the torque M as follows: M = Kcpt and CPt is the
twisting angle for the torsion bar. When the value of t* is commensurable
with the characteristic relaxation time e, the effects of interactions of
elastic oscillations of torsion and viscoelastic effects in polymeric liquids
can be significant. 167 This effect is estimated experimentally by the ratio
CPt/CPp where CPp is the angle of the plate's turning. Though this effect can
be almost removed when using the more rigid torsions, these require
using much more sensitive recording equipment. Similar effects of torsion
bars and diaphragms may occur in measurements of time-dependent
normal stresses.
(d) Inertia and friction effects of instrument fittings

Sometimes it is necessary to take into account the inertia effects of torsion


bars and of some fittings mounted thereon (the cone, a part of a
displacement transducer, etc.), and even their air friction.168 Usually the
last two factors are allowed for in dynamic high frequency tests.
The inertia effects of a torsion bar with fittings and its air friction is
negligible if
UOJ/K 4,1

(4.21)

where OJ is the frequency of oscillations, I is the inertia moment of the


torsion bar with the fittings, U is the coefficient of viscous air resistance
for the torsion bar with the fittings, determined experimentally by
decaying free oscillations, and Mb is the moment on the torsion bar
created by air.

Rheology of Viscoelastic Liquids

89

In experiments under given moment M, the frictions in bearings and


potentiometer as well as the inertia of moving parts, should not exert an
appreciable effect on the movement of weights (see Fig, 4.5). Also, in the
experiments with one-dimensional retardation (see Fig. 4.5(b)), the effect
of the disk inertia is insignificant if
(4.22)
where I is the inertia moment of the disk, e is the relaxation time, 1'/ is the
Newtonian viscosity, s is the ratio of relaxation to retadation times
(0 < s < 1), a is the angle between the cone and plate, and R is the radius of
disk. Inequality (4.22) is obtained by analysis of the experimental scheme
shown in Fig. 4.5(b), using the linear description of the retardation
process by eqn (1.99).
(e) Dissipative heating

Dissipative heating may prove to be important in the case of intense


shearing. A rough upper estimation for temperature increment i1.T within
the time interval t of the experiment can be given assuming the process to
be adiabatic and the liquid Newtonian,
i1.T<

~ft y2(t') dt'


pCp

(4.23)

where cp is the heat capacity at constant pressure. More accurate


evaluations of dissipative heat are given in Chapter 9. The negligence of
the dissipative heating effect on flows of elastic liquids can be verified by
the independence of experimental results from the gap width, because the
effect decreases with the decrease in the gap between thermostatted
surfaces of a rheometer. When dissipative effects are very high, they can
even change the character of flow, e.g. a steady flow may not exist in
shearing under constant tangential stress (see Chapter 9).
(f) Wall slip

Wall slip may also happen under intense deformations applied to a


polymer melt. These can result in failure of adherence of a polymer to,
and its detachment from the wall. In this case, it is natural to assume that
the wall slip of a viscoelastic liquid is similar to the friction of a
crosslinked rubber (see, e.g. Refs 169-171) whose friction force does not
depend on pressure when the latter is high enough, and decreases with
the increase in slip velocity (see Chapter 11). Thus the flow of an elastic
liquid in a cone-plate rheometer with a fixed shear rate O/a, when the

90

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

detachment happens, begins depending on the geometry of instrument.


Sometimes the detachment of liquid from the wall can drastically change
the very nature of the flow, e.g. the steady flow with the wall slip may not
even not exist when Q/oc = constant. 169
(g) Effect of surface tension on normal stresses

The surface tension acting at the free edge surface may appreciably affect
the value of thrust F represented by eqn (4.14) and approximately by eqn
(4.15). It may also change the normal stress distribution obtained by local
measurements, as shown in Fig. 4.3. The easy evaluation of this effect can
be made on the basis of static meniscus:
arrlR ~ 2a* /(ocR)

(4.24)

where a* is the coefficient of surface tension and ocR is the gap width at
the edge. It should be noted that more accurate estimations of surface
tension are difficult since the formulae for static meniscus could not be
applicable for motion. In order to eliminate the effect of surface tension,
the edge of the operating unit is filled up with the liquid tested so that not
only the gap between the cone and disk but also the cup (12) are filled
with liquid (see Fig. 4.3).
(h) The effects of geometry inaccuracy on the measurements of
stresses in cone-plate instruments

Due to errors in setting the gap between the cone and plate, the deflection
of torsion bar (8) (Fig. 4.3) and even because of the deformations of the
body, the cone can be displaced relative to the disk which increases the
gap between these. Figure 4.6 172 demonstrates how the gap between the
cone top and plate, b, influences the results of stress measurements,
especially those for the normal force. This factor is of considerably less
importance when measuring the tangential stresses. To eliminate this
effect, it is possible to use torsion bars with measuring schemes compensating for the deflection 168 or to raise the rigidity of the system with an
increase in the accuracy of measurements. 173 The distortions of alignment and concentricity of the cone and plate may also highly influence
the measurements of normal stresses. 168
(i) The irifluence of defects in compensating schemes on
measurements of normal stresses

Movement of the measuring bar during transitional processes results in


variations of the gap between the cone and plate. In order to reduce the

91

Rheology of Viscoelastic Liquids

86~-l- _

2.8 BW6+- _
88 17. 6 26.4 J5.2 !(mkm)
10 20 30
Ito
x100 %

I
fO

+_
I

20

30

1'U 2"'.4-

It

Fig. 4.6. The effect of a cone's axial displacement on the values of (a) first
normal stress difference and (b) shear stress. LOPE melt, T = 140C.

errors due to gap variations, a compensating scheme for estimating the


thrust F is commonly used (see Fig. 4.3). In this scheme, a displacement
transducer (lO) is usually mounted under the bar (9) whose rigidity is
small. A signal of the value of deflection is received by the device (11)
which displaces the free end of the bar until the deflection is not
compensated. Usually very small gaps (b~ lO2,um) between the cone and
plate are employed to prevent the contact of the cone top with the plane.
Note that in the case of non-steady measurements, such a scheme with
large viscosities (1/ > 105 Pa s -1) and the above-mentioned small values of
b results in large errors. 172 ,174,175 Under conditions y= constant or
(J'12 = constant, the time when, in this interacting liquid-device system, a
steady regime is reached, may be raised by several orders of magnitude
while the variable component of normal stresses is not registered under
high-frequency oscillations at all. This is due to the existence of small
displacements to which measuring devices are not sensitive. In order to
pass through a 'threshold distance' Li in such a system, a time interval
t ~ 1/SLi/(bF) is required (S is the area of disk) which may be significant
when the values of b are small but 1/ is large. When using such a scheme,
even the attainment of a stationary situation with viscosities 1/ > 107 Pa S-l
is very problematic owing to the slow performance of the monitoring
system. 168 To eliminate these undesirable phenomena the gap b might be

92

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

increased by handling, for example, the disk-disk unit without changing


the nature of flow.
(j) M echano-degradation and air bubbles
During long time experiments, mechano-degradation may occur, which is
sometimes followed by polymer foaming. This effect results usually in
very slow variations in registered indications.
With the cone and plate coming closer to each other, and particularly
when the tested liquid is of low viscosity, air bubbles sometimes form at
the top of the cone. These effects can significantly influence the resulting
measurements.
Concluding this section we shall note that in performing the experiments, some efforts are needed to find a proper cone-plate unit, tension
bars and fittings so that the mentioned undesirable effects can be
negligible in the region of the experimental window. Most of the errors
and effects considered here also appear in experiments made on other
operating units, such as cylinder-cylinder or disk-disk units.

4.3 DISK-DISK RHEOMETER


A disk-disk rheometer is shown in Fig. 4.7. The test liquid (1) is placed in
the gap between parallel and coaxial disks (2 and 3). Here one disk (2) is
steady and the other disk (3) is rotating with an angular velocity Q which
is usually constant. Under isothermal conditions, the only non-zero
component of the strain rate tensor is:
(4.25)

where H is the distance between the disks, {r, cp, z} are the cylindrical
coordinates; the z-axis being directed along the axis of rotation. The
stress tensor has the following non-zero components: l1m l1cpcp, l1 zz and l1 rz
depending only on the radius r and generally on time t. As seen, the
simple shear in flows between two disks is inhomogeneous.
Experiments with flow between coaxial disks are commonly carried out
using the same measuring apparatus and torsion bars as with the
cone-plate units. Measured here, are the torque M and thrust F which
are represented as follows:
M = 2n

LR

l1

cpAr)r 2 dr

(4.26)

93

Rheology of Viscoelastic Liquids

F= -2n

f:

azz(r)r dr=n

tR

[adr) + a2{r)Jr dr

(4.27)

where a 1 = a",,,, - a rr and a 2 = a rr - a zz are the first and second normal


stress differences, and R is the radius of disks. Expression (4.27) for the
thrust follows from the inertialess equations of momentum balance (stress
equilibrium), allowing for stresses to be independent from the coordinates
z and <p.

.......-.Q

Fig. 4.7. Disk-disk rheometer: 1,


tested liquid; 2, unmovable disk; 3,
rotating disk.

Fig. 4.8. Capillary rheometer: 1,


tested liquid; 2, reservoir; 3, piston; 4,
capillary.

The basic effects and errors in measuring the values of torque M


and thrust F are the same as for the flow in cone-plate devices. There
is an advantage, however, in using the disk-disk unit over the coneplate version in that there is the opportunity to vary appreciably the
gap H between the disks without noticeably changing the shear rate,
i.e. the character of flow. This makes it feasible to measure, with the
help of a compensation scheme, the integral characteristics of normal
stresses, F, under non-stationary flow conditions. The main disadvantage of this device is the impossibility of measuring the local
normal stresses by measuring the thrust, which is due to the shear rate
inhomogeneity. The correctness of the results obtained is controlled
by their independence of a change in H under fixed integral shear rate
OR/H.

94

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The isothermal flows of elastic liquids between the disks are examined
in Chapter 5; the flows with dissipative heating are considered in
Chapter 9.
4.4 CAPILLARY RHEOMETRY
4.4.1 Basic Elements and Operational Procedure
A sketch of a capillary rheometer employed in tests of polymeric melts and
concentrated solutions is displayed in Fig. 4.8. Here an elastic liquid (1) is
pushed away from the reservoir ('bomb') (2) through a capillary (4) by a
piston (3). Usually the bomb greatly exceeds the capillary in size. The
cylindrical capillary has the length L and diameter Do = 2Ro where usually
L/Do ~ 1. Sometimes, especially when flow visualization is required, a flat
capillary is used with Do being the gap width between the planes and
L/Do ~ 1. The bomb and capillary are thermostatted. Various technical
models of capillary rheometers have been discussed in the literature (see, e.g.
Refs 176-178). Usually the motion of the piston in the bomb is considered as
very slow, thus in a good approximation one can assume that the stressed
state in the bomb is close to the hydrostatic equilibrium.
There are two major types of capillary flow: pressure controlled and
rate controlled. In the pressure-controlled flow, the pressure p in the
bomb is given and is usually a constant, whereas the flow rate Q is
measured. In the rate-controlled capillary flow, the velocity of the piston
is given and is usually constant, which for an incompressible liquid
ensures the flow rate Q to be constant; the pressure p in the bomb being
the measurable variable.
Capillary rheometers are commonly employed to measure the flow
curves in higher regions of shear rates when the cone-plate and disk-disk
rheometers are very difficult to use due to intense chipping of a tested
liquid out of the operation gap. Experiments with capillary rheometers
require usually much more polymer than those using rotation devices.
Sometimes this circumstance is important.
4.4.2 Determining Flow Curves in Isothermal Case
A steady flow in the capillary is usually considered under the assumption
that the capillary has infinite length. In this case, there is only one
component of velocity, Vz dependent on radius r and, perhaps on time t,
while the shear stress, {1m in the capillary is given by:
{1rz=rf /2,

f=f(t)=Ap/l

(4.28)

Rheology of Viscoelastic Liquids

95

where Ap is the pressure drop in the capilary and 1(1 < L) is a characteristic capillary length whose value is discussed below. Here we use the
cylindrical coordinates {r, qJ, z}; the z-axis being directed along the
capillary axis. There are also normal stresses, appearing usually in the
capillary flows of elastic liquids, which usually are dynamically insignificant in developed flows. This flow is known as 'developed' flow, meaning
that it is independent of longitudinal coordinate z. A more detailed
description of capillary flow is given in Chapter 8. Equation (4.28) is
easily derived from the equations of stress equilibrium. It shows that the
simple shear flow in capillary is inhomogeneous.
In order to determine the flow curve, i.e. the dependence of the shear
rate Y on the shear stress (T = (Trz in a steady flow, we use the familiar
relation (see, e.g. Ref. 101):
.
4Q
Yn= R 3 '

n
where Yn is the shear rate at the capillary wall defined without allowance
for non-Newtonian properties of the liquid, and (Tw= (Trz(Ro) is the shear
stress at the capillary wall. The values of Yn and Ap are measured during
the capillary experiments.
When, in a rough approximation, the effects of capillary entry and exit
on the flow are not taken in the account, the pressure drop Ap is
represented as the difference between pressures in the bomb and at the
capillary exit, while 1= L, where L is the capillary length. Experimental
dependences Yn((Tw) can be approximated within a large region of (Tw
variations, by using the power 'law':
0:=(3a+ 1)j(4a)
In the case of Newtonian liquids, a =
solutions and melts, a < 1.

0: =

(4.30)

1, while for the polymer

4.4.3 Entry and Exit Corrections in Capillary Flow

Now let us consider in more detail the flow of elastic liquids in a capillary
rheometer. When the flow is intense enough, before entering the capillary,
a very remarkable flow in the reservoir also appears.176.179-181
As shown in Fig. 4.8, the main part of this flow forms an 'entrance cone'
with (usually) slowly rotating vortices outside. In the capillary itself, the
flow close to the developed one occurs only at the distance 11 downstream
from the entrance. For low Reynolds flows of Newtonian liquids, 11 ~ R o ,
but for elastic liquids, even when Re ~ 1, the values of 11 are much

96

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

larger.180.181 Also, the exit flow is very different from the developed flow
and is widespread on the length 12 upstream. In the entrance and exit
regions, some complicated, usually two- or three-dimensional flows
happen, related to the rearrangement of velocity profiles. It is worth
mentioning that for elastic liquids, the length of the entrance zone II is
greater (sometimes, considerably greater) than the length of exit zone
I2' 181
In order to reduce this complicated behaviour to a one-dimensional
situation, Bagley179 suggested the following procedure usually known as
'Bagley's correction'. It is assumed that the terms (Jw and Yn((Jw) in eqn
(4.29) still hold; however, with magnitudes ,1.P and I presented in definition (Jw, now changed as follows:
(4.31)

Here the term PI reflects the total pressure losses including the precapillary flow, as well as the entrance and exit regions of capillary; the
quantity ,1.Pe is the measured pressure drop; II and 12 are the lengths of
entrance and exit zones. Though the values Pi> II and 12 are unknown, we
can estimate these from experiments with various capillary lengths Land
the same radius Ro under condition Q= constant.
4.4.4 Dissipative Heating
Unlike the flow in rotary rheometers, discussed above, even a significant
dissipative heating in capillary flows under steady conditions ,1.P = constant or Q= constant, does not violate the stationarity of flow. It could be
shown (see, e.g. Ref. 178) that the dissipative heating in the flow of a
Newtonian liquid is negligible, if
mQ,1.p/(16nqKL)~

(4.32)

where q is the mechanical heat equivalent, K is the thermal conductivity


and m is the temperature viscosity coefficient when the temperature
dependence of the viscosity is represented as: 11 = 110 exp[ - m(T - To)].
The estimation (4.32) is also valid for viscoelastic liquids. It should be
mentioned that the flow curves in steady flows can be determined even
when the heating effects are great. 178
4.4.5 Evaluation of Wall Slip
When Deborah numbers in capillary flows of polymers are high, wall slip
may occur due to the failure of adherence of a polymer to the wall. But a
steady state flow may still exist. In this case, the methods of capillary

Rheology of Viscoelastic Liquids

97

rheometry are useful to determine the stationary slip velocity. In the


following it is assumed that (i) the capillary length L is large enough to
neglect the Bagley corrections and consider the flow of polymer as
completely developed and (ii) that the slip velocity, u, depends only on the
wall shear stress, O"w' Then in this case, the magnitude Yn in eqn (4.29) is
changed for the following:
Yn=4Q/(nR6)-u/R o

(4.33)

Equation (4.33) allows us to make measurements of the flow curve in


the steady-state capillary flows with wall Slip.182.183 With a fixed value of
wall shear stress, o"w, one can determine the dependence 4Q/(nR6) on l/Ro
which as shown in experiments, is linear. Then the value of slip velocity, u
is determined by extrapolation of the dependence of l/Ro-+oo.
Some direct measurements of wall slip in capillary flows, which are
usually accompanied by oscillations, were carried out using different
methods which are briefly discussed in Chapter 11.
4.5 EXPERIMENTAL METHODS IN THE UNIFORM
EXTENSION OF LIQUID POLYMERS

As shown below (see Chapters 6 and 7), elastic liquids behave quite
differently in nonlinear regions of stretching and simple shear deformation. Additionally, in extensional flows there is no poorly controlled
interaction of a flowing polymer and the wall. This is a favourable feature
of extensional deformations as compared to simple shearing.
4.5.1 Basic Relations
We will consider here a uniform inertialess extension under isothermal
conditions. The velocity components in the Cartesian coordinate system
{Xt.X2,X3} are represented as follows:

(4.34)

(i= 1, 2,3),

where the second relation in eqns (4.34) reflects the condition of incompressibility. In the following, the axis Xl is directed along the main axis of
extension. The strain rate and vortex tensors are of the form:

m=O
where in general, the values Ki depend on time.

(4.35)

98

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Commonly studied is a uniaxial extension of a specimen with cylindrical form, which is easier to carry out than other types of extension. The
particular relation for the strain rate tensor in this case, is of the form:

e=K diag{l,

-t -!}

(4.36)

Also, in the cylindrical coordinate system {r, qJ, z} with the z-axis directed
along the extension, the velocity components are:
Vz =

K(t)Z,

Vr

-K(t)r/2,

(4.37)

In uniaxial extension, the diameter of specimen does not depend on


spatial coordinates and changes only with time t. The uniform extension
is usually performed either by assigning a certain 'law of motion' (or
loading) at the mobile end of a specimen, or a certain velocity Vz = K(t)L
(or loading) in a fixed place z = L.
If the effect of surface tension is negligible, there is only one component
of stress tensor, (J zz == (J, in the uniaxial extension, which does not depend
on the spatial coordinates; the stress being related to the stretching force
F and the radius of specimen R(t) by the relation: (J=F/nR2. The surface
tension is insignificant if y/R ~ (J, where y is the coefficient of surface
tension.
The relative elongation or stretching ratio of an incompressible specimen is expressed as follows:
(4.38)

where 10 and Ro are the length and radius of the specimen at the initial
instant, and 1 and R at an arbitrary instant of extension (see Fig. 4.9). In
the case when the length of a specimen under extension is fixed, the

Fig. 4.9.

A specimen in initial, deformed and unloaded states.

Rheology of Viscoelastic Liquids

99

second relation in eqn (4.38) is used to determine the quantity . If the


dependence (t) is known, the strain rate K(t) is determined by:
K=dln/dt

(4.39)

This relation results from the general kinematic equation (A.40) under
conditions that w=O and C- 1=diag{2,e-1,-1}, where C- 1 is the
Finger total strain tensor.
In addition to the complete strain (t), it is convenient to determine in
elongation tests, the recoverable (or highly elastic) strain,
&(t) = I(t)/Ir(t) =

R; (t)/R2(t)

(4.40)

Here Ir and Rr are the limit values at t ....Ht) of length and radius of
a specimen after its instantaneous unloading (see Fig. 4.9). This is a
typical retardation experiment where the contraction of the specimen
occurs due to an elastic energy accumulated during extension of a
polymeric liquid.
The following quantity is taken in experiments as the magnitude of
irreversible longitudinal strain: 184 ,185
In/3 = ln -Ina

(4.41)

Then the irreversible longitudinal strain rate, ep is defined as following: 186,187


ep = d In/3/dt = K- d In&/dt

(4.42)

Equations (4.41) and (4.42) can also be represented in the tensor form if
both the sides of each equation are multiplied by a constant tensor
P=diag{l, -!,
Let us also present the useful kinematic formulae 187 which relate the quantities I(t), R(t), K(t) and the longitudinal
velocity of mobile end, v(t):

-no

In addition to these experiments, the regime of stress relaxation is also


considered where, after instantaneously stopping the stretching process,
the length of a specimen is fixed and the stress value is decaying with
time.

100

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The most common regimes of uniform extension are: under constant


velocity of mobile end, V,186,187 under constant strain rate, K,188-192
under constant force, F 184 ,193 and under constant stress. 194 ,195
4.5.2 Schemes of Instruments
Figure 4.10 sketches the simplest device with a varying base. The
specimen (1) is attached to the clamps (2 and 4). The immovable clamp
(2) is connected with a device (3) measuring the stretching force, F.
The mobile clamp (4) moves with a given speed. To compensate the
weight and for thermostatting, the specimen is placed into the bath (5)
with a low viscous liquid whose density is close to that of the tested
polymeric liquid.
/

Fig. 4.10. Extension of a various length cylindrical specimen: 1, specimen; 2,


fixed clamp; 3, device for measuring force; 4, movable clamp; 5, thermos tatting
bath; 6, float; 7, thread; 8, roller; 9, load.

If assumed that at the mobile clamp the velocity v = Vo = constant, the


strain rate K is given by:
(4.44)

where 10 is the initial length of the specimen.


When K = Ko = constant, the movable end of the specimen travels
according to:
l(t) = 10 exp(Ko t)

(4.45)

In the cases when a=ao=constant or F=Fo=constant, a feedback


system is commonly used where the signals from a force transducer
control the motion of movable clamp. Its displacement is assigned so that
under a given constant stress the measured force F=ao Vo/l(t) (Vo is the

Rheology of Viscoelastic Liquids

101

constant volume of specimen), or F = constant when the constant force is


given. In the latter case, the stress is calculated as:
a(t) = F o(t)/Vo = (t)F o/(nR6)

(4.46)

The simplest way to perform an elongational experiment under F =


constant is to use a constant loading 193 as shown in Fig. 4.10. Here the
clamp (4) is mounted with a float (6), connected by thread (7) via the roller
(8) with a load (9). If the roller has a certain profile,194 the same method
can also be used for imposing the condition a = constant.
The condition v = constant can be simply achieved by winding the thread
tied to the float, on a roller rotating with a constant angular velocity.
Consider now the schemes of elongational devices with a constant base,
represented in Fig. 4.11. Figure 4.11(a) shows the scheme of Meissner's
device. 189 ,190 Here a specimen (1) is placed between two pairs of toothed
rollers (2 and 3) rotating with a given angular speed in the directions shown
by arrows, and being at a distance L from each other. The rollers (2) are
rigidly fixed with the device (4) measuring the longitudinal force F.
In the case shown in Fig. 4.11(b),192 the specimen (1) is fixed with one side
to the unmovable clamp (2) which is rigidly connected to the device (4)
measuring the stretching force. Another side of the specimen is fastened to
the drum (3) with radius R 1 , rotating with a given angular speed. The drum
can be shifted in an axial direction for winding a multiple specimen. Also,
there is another arrangement of the device where the right-hand side of the
specimen is pulled in the gap between gears, as shown in Fig. 4.11(a). To
provide the regime K = constant the rollers shown in Fig. 4.11 have to rotate
with a constant speed. In this case, the strain rate is:
(4.47)

where 0 1 and O 2 are the angular velocities of the roller's rotation. For the
diagram shown in Fig. 4.11(b), O 2 =0.

(a)

Fig. 4.11. Uniform extension of polymer melt with a constant base. (a) Meissner
device: 1, specimen; 2, 3, rollers; 4, device for measuring force. (b) A device with
one fixed end and one roller: 1, specimen; 2, fixed clamp; 3, roller; 4, device for
measuring force.

102

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In experiments with a given stress or force, the variations of dimensions


of specimen (l(t) for the schemes with a varying base, and R(t) with a
constant one) are measured in time. When the kinematics of extension is
given, the force F(t) is measured and, thereby, the stress O'(t).
In all the devices described, a stretched specimen can be cut off,
practically instantaneously, from the moving parts. After that, the retardation process occurs in freely floating specimens, which due to eqn (4.40)
allows us to determine the recoverable strain &.
Using the devices considered, one can also perform the stress relaxation experiments. For instance, for the scheme shown in Fig. 4.10, the
relaxation after a given elongation ratio is easily achieved in an obstacle
in the form of a fork is mounted at a certain distance from the unmovable
clamp.193 The relaxation starts when the float, in the course of its motion,
clings against the fork.
In the experiments with a given constant velocity of movable end, due
to eqn (4.44), K-+O when t-+CIJ which also give an opportunity to study
the linear region of deformation.
The regimes K = constant and 0' = constant are favoured for the simplicity of theoretical treatment of experiments. Under such deformations,
elastic liquids usually attain a steady flow, if the level of stresses is not too
high. The regimes of simple elongation under assigned constant force give
the highest intensity of deformations of those considered. It is easy to
achieve experimentally and it is very close to the real industrial processes
for forming films and filaments.

4.5.3 Methodological Aspects

From a methodological viewpoint, the experiments with simple elongation of polymer melts are much more complicated as compared with
those for the solids. These difficulties are due to: (i) large strains
developing usually in elastic liquids where the elongation ratio can reach
up to 12; (ii) troublesome operations with decomposition of the total
strain into reversible and irreversible parts; (iii) the necessity of measuring
fairly small forces which are often under 1O- 2 N; and (iv) difficulties with
the fixation of a liquid polymer in clamps.
Cylindrical samples are usually made of fairly viscous polymers with
Newtonian viscosities 1]? 10 5 _10 6 Pa s -1, so that they might not spread
about in course of their fastening and subsequent stretching. Samples
of melted polymers are made by either extrusion from a capillary into
equidense low-viscous liquid 186 ,187 or by pressure. 188 ,190,191 When
the melting point of a polymer is lower than the room temperature, as

Rheology of Viscoelastic Liquids

103

e.g. for polyisobutylene, it is convenient to prepare the specimen by


rolling it between two plates until the required diameter is achieved. The
diameter is assigned by using balls of a certain size. The samples are
usually prepared with the initial radius Ro ~ 0.2-0.3 cm and the ratio
lo/Ro> 10.
Samples are fixed to other units, such a system of force measurement,
float, etc., by means of adhesive tape, glue or bushing (shown in Fig. 4.12).
A conical bushing (2) is put on the collar (3) connected with the
immovable part of the device or float. In the conical bushing, the end of
the specimen (1) is fixed by the same polymer when the space between
polymer and bushing has been filled.
During the tests, the specimen is in the bath (with a thermostatting low
viscous liquid whose density is close to that of the polymer), or the
polymer is floating over the liquid.
Before testing, a specimen placed in the bath is kept for some time in an
unstrained state. This is done to remove residual stresses occurring in
sample after its preparation or heating. The time tl needed for the removal of residual stresses is proportional to the characteristic relaxation
time O. The time for heating the sample in the bath, t2 is proportional to
R6/fi where fi is the thermal diffusivity.
f

Fig. 4.12. A scheme of fixing of a specimen by


bushing: 1, specimen; 2, conical bushing; 3, collar.

4.5.4 Defects in Schemes of Devices and Errors in Measurements


An evident defect of an elongation device with a varying base is the
restricted travel of the movable clamp, and hence, the strain limitations.
In the case of devices with a constant base, a defect of the four-roller
scheme shown in Fig. 4.11(a) is the difficulty in assembling the rotating
rollers and the force measuring device. When one end of the specimen is
fixed (as shown in Fig. 4.1 1(b)) a non-uniform region occurs at the fixed
end which increases with the increase in strain. This can noticeably affect
the results of the measurements if the length of this region is not much
less than the base length L.

104

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

It should also be noted that when a specimen is extended by a pair of


rollers, a partial extrusion of polymer at its entry into the gap between
rollers may occur, which results in errors related to inequality
v(L) 0;6 QR. 192 When extended by one roller, an effective slippage of
the specimen over the roller can happen, which results in the same
error. The check of such errors as well as the observations of inhomogeneous regions most common with small radii of a stretched
specimen, may be exercised by means of photographing and comparing
the calculated diameter of the specimen with the actual one.
When stretching a specimen with a load (see Fig. 4.10), self-oscillations
may also arise due to the interaction of the thread friction and the
elasticity of the specimen. 196
The deformation of a dynamometer is negligible when Gt~Ge, where
Gt and Ge are the elastic modulus of the measuring bar and the elasticity
modulus of the specimen, respectively. In this case, the results of tests
should be the same when the bars are changed.
The effect of dissipative heating can be roughly evaluated when
assuming the extension to be adiabatic:
(4.48)

where d T is the increase in temperature due to the heating, rJ is the


Newtonian velocity of an elastic liquid, K is the strain rate, p is the
density, cp is the heat capacity at constant pressure, and t is the actual
time of experiment.
When measuring the recovery strain in retardation experiments, the
time needed for the contraction of the specimen is assumed to be of
the order of the relaxation time e. The sample may also contract
under the action of surface tension which over a certain time imparts
the specimen with a spherical form (see Fig. 4.13). The effect of surface tension can be measured by a comparison of contractions of
strained and unstrained specimens of the same initial sizes. 186 Usually,
the surface tension driven contraction is much slower for polymer
melts than that due to the retardation and therefore, it can be disregarded.
In order to accelerate the retardation, sometimes the specimen is
heated. This may result, however, in an error of the residual length lr
measured 197 because of a partial loss of elastic energy for relaxation
processes owing to non-isothermality across the specimen.

Rheology of Viscoelastic Liquids

Fig. 4.13.

105

A cylinder of polyisobutylene contracting under action of surface


tension.

4.6 EXPERIMENTAL METHODS IN NON-UNIFORM


EXTENSION OF LIQUID POLYMERS
4.6.1 Weakly Non-uniform Extension
It is convenient to measure the rheological properties of relatively
low-viscosity polymeric liquids under non-uniform extension. Also, this

106

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

type o.f experiment is used when simulating so.me techno.lo.gical processes.


Co.mmo.nly the weakly no.n-unifo.rm extensio.n is used. In this case
(depicted in Fig. 4.14), the radius R o.f the specimen slo.wly changes alo.ng
the lo.ngitudinal co.o.rdinate z, so. that JdR/dz ~ 1; the velo.city co.mpo.nents in the cylindrical co.o.rdinates being the fo.llo.wing:
J

Vz~VI +

J:

K@

d~,

(4.49)

Vr ~ RK(Z)/2,

where VI is the lo.ngitudinal velo.city at z = 0, and k(z) is the strain rate.


When the surface tensio.n, weight and inertia effects are insignificant, the

~L.
0

Fig. 4.14.

'lit

~T

--

V'g

Profile of specimen under weakly non-uniform extension.

lo.ngitudinal stress co.mpo.nent u(z) is related to. the co.nstant stretching


fo.rce F as fo.llo.ws:
(4.50)
If the dependence R(z) is determined fro.m, e.g. pho.to.graphy, and if the
flo.w rate Q (independent o.f z), is also. kno.wn, the velo.city distributio.n
alo.ng a specimen is given by:

(4.51)
In no.n-unifo.rm extensio.n, each sectio.n o.f the specimen, perpendicular
to. the elo.ngatio.n axis z, travels alo.ng this axis in time as in the case o.f
unifo.rm extensio.n with the same fo.rce F. The fo.rmulae fo.r the transformation o.f the data o.btained in the unifo.rm extensio.n experiments into.
tho.se fo.r a no.n-unifo.rm case are given in Chapter 10.

4.6.2 Schemes of Devices for the Drawing of Elastic Liquids


with a Free Surface
The scheme o.f a device fo.r the drawing o.f elastic liquids with a free
surface is sketched in Fig. 4.15. An elastic liquid (1) under study is
statio.nary drawn fro.m the reservo.ir (2) by means o.f a rotating drum (4),
and flo.ws back alo.ng the gro.o.ve (5) into. reservo.ir; at the first mo.ment,

Rheology of Viscoelastic Liquids

107

2
(b)

(a) ~='77777'77777=="fFig. 4.15. Drawing of a polymer solution with free surface; (a) an experimental
device, (b) groves for jet's return: 1, stretching jet; 2, reservoir; 3, wire hook for
measuring stretching force; 4, rotating drum; 5, groove; 6, springs.

the liquid is pulled up to the rotating drum by a glass stick. Thus, the
level of liquid in the reservoir is maintained constant at steady drawing.
In order to keep the liquid steady on the drum, the latter is manufactured
with a small conicity in its middle. The groove (5) which is an acrylic plate
conforms to the contour of the drum and is pressed against it with a
spring (6), as shown in Fig. 4.15(b). In experiments where the vertical
stretching force F is measured, a withdrawing polymeric jet, before
getting on the drum, passes through a wire hook (3). The hook is
positioned so that the axis of the part of jet coming from the reservoir
might make a right angle with the axis of jet's part going to the drum.
The hook is rigidly fastened to a transducer measuring the force F.
Moreover, the flow rate in the jet can be found by a measuring glass
placed under the running jet. The distance between the free surface of the
liquid and the drum can be varied. The profiles of jet, i.e. the dependence
of its radius on the vertical coordinate, can be determined from photographs.
It turns out that not all liquids can be drawn in such a way. An ability
to be stretched from a free surface is called spinnability and the scheme of
the device considered is intended for studying this feature of polymer
solutions, which is of great importance in the fabrication of fibres and
films.
In Refs 22 and 198, stretching from the free surface was performed with
the help of an evacuated tube. At the first moment, the tube is lowered into
the reservoir, then it is raised with a liquid column inside over the free
surface. The reservoir is placed on scales to measure the stretching force.

108

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

4.6.3 Drawing Out of a Capillary


Drawing out of a capillary can be done with a device similar to that
sketched in Fig. 4.16. 199 The jet (2) of a polymer solution is extruded from
the capillary (1) at a constant flow rate and a further is drawn by a
conveyor (3). The capillary also serves as a bar whose small travel is
recorded by a transducer (4) to measure the stretching force. A profile of
the extrudate is photographed and the flow rate Q is also measured.

2
+

Fig. 4.16. Drawing of a polymer solution from


capillary by means of conveyor: 1, capillary; 2,
polymer jet; 3, conveyor; 4, transducer for measuring stretching force.

Figure 4.17 demonstrates steady drawing under a constant load (6), of


a polymer melt extruded from a capillary (1), commonly at a constant
pressure. The load is connected with the extrudate by a thread (4) passed
over the rotating roller (5). The extrudate is fastened to the foat (3) by
means of a conic bushing as in the case of uniform extension (Fig. 4.10).
To compensate for the weight and for thermostatting, the extrudate is
kept floating in the bath tube (7). The jet's profile is photographed by a
camera (8). This kind of experiment, along with its theoretical treatment,
is considered in Chapter 10.
Some other schemes of devices for isothermal non-uniform drawing
can be found in Ref. 200 and these for non-isothermal conditions
common in the industry, are given in Refs 201 and 202.

s
6

Fig. 4.17. Drawing of polymer extrudate from capillary under constant load: 1,
capillary; 2, extrudate; 3, float; 4, thread; 5, roller; 8, load; 7, bath; 8, camera.

Rheology of Viscoelastic Liquids

109

4.6.4 Some Methodological Observations


All the schemes for non-uniform extension considered have a significant
limitation which does not allow us to carry out fairly accurate quantitative rheological studies. Though from these experiments, the stress (J and
strain rate Ii; are known functions of longitudinal coordinate z, with no
theory at our disposal, or experimental data on the uniform situation, we
cannot evaluate the effect of the shear deformation pre-history in a
capillary flow on the subsequent extension of a free extrudate. The same
is true for the drawing from a free surface where there is a strongly
non-uniform deformation of a polymeric liquid at the free surface with
heavy effects of surface tension (meniscus).
Since the polymer jets can slip over the drum and conveyor, it is not
enough to know their speeds; it is also necessary to measure the flow rate
for a correct evaluation of the strain rate.
The possible detachment of polymer from the capillary wall cannot be
ignored and to check this visually the capillary is made transparent. The
detachment can be also detected when the extrudate's diameter measured
at the non-transparent end of capillary by photography, is of the same
diameter as the diameter of orifice.
For the drawing of polymeric liquids from a free surface (Fig. 4.15), one
has to check the possible effect of the hook (5) for measuring the force on
the profile of drawn jet. Usually the effect is negligible if the jet profile is
independent of whether the hook is employed or not. The infuence of the
conveyor or drums in Figs 4.16 and 4.17 is evaluated in the same way.
The profile of a fixed length under a fixed flow rate should not vary when
displacing tlw drum (conveyor) relative to the free surface (capillary).

4.7 METHODS OF FLOW BIREFRINGENCE


In addition to the usual methods of rheometry, some optical methods,
especially flow birefringence are widely used nowadays. The great advantage of the methods is that these do not disturb the flow patterns and can
give valuable information, particularly about the normal stresses and
stress distribution in complicated flows of polymeric liquids. But these
methods need more complicated experimental procedures and instruments.
In isotropic transparent media, the speed of light propagation, v, is
different from its value, c, in vacuum. The refractive index in these media
is defined as n = c/v. On a molecular level, the phenomenon of refraction

110

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

is explained by the polarization of molecules under action of electric field.


If the electric field E is not too strong, the polarization vector P is
proportional to E, i.e. P = 0( E where 0( is the polarizability coefficient. The

refractive index is related to polarizability by the familiar LorentzLorenz formula:


n 2 -1

4n

n2 +2

--=-0(

(4.52)

These relations are changed in transparent anisotropic media, where


the light propagates with different speeds in different directions. Instead
of the refractive index, the optical anisotropy is characterized here by the
refractive index tensor n. This is a second-rank symmetric tensor whose
principal axes ni are related to the principal optical axes of a specimen,
along which the light propagates with different, in general, velocities
Vi = c/ni' The optical anisotropy can be characterized by the symmetric
second-rank polarizability tensor oc. The relation between tensors nand oc
is vague.
Apart from intrinsic anisotropic media like crystals, there is a vast
number of isotropic, usually amorphous, materials which gain the optical
anisotropy only under load. This stress-induced optical anisotropy was
first discovered at the beginning of the last century by Brewster 203 who
also suggested using the phenomenon in the study of stresses in glasses.
Later, Neumann, Maxwell, Wertheim, and many others proposed some
phenomenological relations between stress tensor a (or infinitesimal
strain tensor 8) and refractive index tensor n for amorphous solids. Of
these, the only Maxwell linear relation: 204
n-

(j

trn/3 = C(a- (j tra/3)

(4.53a)

where in a wide range of stresses C is constant, remained valid for a vast


class of solids and liquids. Equation (4.53a) is known as the stressoptical law and constant C is called the stress-optical constant. Equation
(4.53a) is usually employed in the principal axes where it takes the
form:
(4.53b)
The differences in the principal values of the refractive index tensor are
usually measured, however small they may be. In the most important
particular case when n2 = n3' the difference dn = (n 1 - n2) is called double
refraction or birefringence.

111

Rheology of Viscoelastic Liquids

The methods of birefringence find a lot of applications in photoelasticity where they are used for stress analyses of models of complicated
engineering constructions (see, e.g. Ref. 205).
In the simplest case, an elastic transparent glassy plate under planar
deformation can produce a birefringent pattern. The scheme of a simple
optical device, plane polariscope, which is used to determine the direction
and intensity of the principal stresses is shown in Fig. 4.18. Here the light
from a monochromatic light source (usually a laser beam) is propagating
through a polarizer with vertical optical axis. The polarized light propagates through the stressed plate (specimen), whose optical axes coincide
with the axes of principal stresses and turned on angle rx relative to the
polarizer. Here the phenomenon of double refraction occurs, which is
then analysed through the interference pattern on an analyser and then
on the screen. The easy optical analysis of this device results in the
expression:
(4.54)

=ndlln/A

where I is the intensity of light, d the thickness of specimen, lln the


birefringence and A the wave length. Equation (4.54) means that the light
does not reach the screen when either sin 2rx = 0 or sin = O. The first case
leads to the occurrence of dark lines on the screen which are called
isoclinics and which allow us to determine the direction of the principle
axes. The second case relates to the occurrence on the screen, of dark

Light
Source

Polarizer
Specimen

Analyser
Screen
Fig. 4.18.

A scheme of plane polariscope.

112

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

fringes of order k, where k is an integer. These give the opportunity to


determine the birefringence dn and, with the use of linear stress-optical
relation (4.53), the difference in principal stresses. Some additional considerations 205 are also used to find the values of components of the stress
tensor. Other, more complicated polariscopes were elaborated and successfully applied to stress analyses in solids and liquids.
In 1943, Kuhn and Griin 206 derived the stress optical relation for
crosslinked rubbers from molecular considerations of polymeric networks. In the Gaussian case, they obtained eqn (4.53) with the stressoptical constant C expressed as follow:
(4.55)
where n is an isotropic part of the refractive index tensor; <Xl and <X2 are
axial and transversal polarizabilities, respectively, in a typical macromolecular link; k is the Boltzmann factor; and T is the Kelvin temperature. In their derivation, the difficulty with the unknown relation between
tensors n and IX was overcome by the assumption that the principal axes
of these tensors are related by the same Lorentz-Lorenz's eqn (4.52) as in
the isotropic case. Equations (4.53) and (4.55) were tested many times and,
in general, a good coincidence with data was found in the neo-Hookean
region of deformations (see, e.g. Chapter 9 in Ref. 16).
In 1956, Lodge 72 extended Kuhn and Griin's theory for polymer melts
and concentrated solutions using the concept of temporary networks.
Filippoff, Wales, Janeschitz-Kriegl and their co-workers, intensively
verified the stress optical relation (4.53) for many liquid polymers. Many
new apparatus were elaborated and applied to study the rheology and
fluid mechanics of elastic liquids. In these instruments, the flow patterns
are used instead of the elastic specimen shown in the optical scheme in
Fig. 4.18. These results are summarized in Ref. 3. A good brief introduction in flow birefringence of polymers can be found in Chapter 6 of Ref. 7.

CHAPTER 5

Theoretical and Experimental Investigation


of Shear Deformations in Elastic
Polymeric Liquids
5.1 INTRODUCTION
In this chapter, the results of theoretical and experimental investigations
that analyse the inertialess shear flows of polymeric liquids under isothermal conditions are presented. The simplest case of shear strains, traditional in studies of liquids, is the flat-plate Couette flow shown in Fig. 4.1.
Other methods of affecting simple shear flows are discussed in Chapter 4.
Under study in this chapter are the flows with either an assigned
motion of the mobile surface (rate-controlled flows) or with a tangential
stress given on it (stress-controlled flows). These flows, which admit exact
solutions, are verified experimentally and occur in various technological
polymer processes.
Rate-controlled flows under study include simple shearing at constant
shear rates, stress relaxation after shearing (including those after a strain
imposed instantaneously), low-amplitude linear and nonlinear steadyshear oscillations and the superposition of low-amplitude oscillations on the
steady-shear flow with both parallel and orthogonal schemes; the latter
being two-dimensional flow. Along with these uniform flows, some nonuniform shearing such as the flow between two parallel coaxial disks (one of
them rotating) are also studied in this chapter. One-dimensional shearing
non-uniform flows in channels and capillaries are considered in Chapter 8.

114

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

For stress-controlled shearing, the uniform shear flow under a given


constant shear stress and the elastic recovery after flow are considered.
This chapter also describes the peculiar phenomena observed in shear
deformations of concentrated polymer solutions and melts. Elastic strains
develop along with flow in these liquids, resulting in the appearance of the
relaxation time(s) which, for the fluids under study, appreciably exceeds the
characteristic time scale of inertial processes. This results in the quick
adjustment of inertia effects to long-time relaxation processes and justifies
the common use of the inertialess approximation. Besides, in all the shear
flows of elastic liquids, normal stresses also exist along with tangential ones,
owing to the development of finite elastic strains in the flow. This is the
Weissenberg effect,21 well known for elastic liquids. This and other
nonlinearities in elastic liquids also result in the fact that two shear flows
superimposed on each other are not independent but highly interactive.
For a general theoretical description of the nonlinear behaviour of polymer liquids in simple shear flows, we use the multimodal viscoelastic constitutive equations (1.71H1.76) with parameter n #- 1 and decreasing dependence of the relaxation times on elastic potentials Wk as shown in eqn (1.59b).
When considering specific problems, we will also use the particular case
n= 1 and y =Oin eqn (1.59b). In all the theoretical treatments, a quantitative
comparison is carried out between the predictions and the data.
In conclusion, a possibility for a wider description of the data by the
specific models38~4o with n #- 1 and y > 0 is discussed using the steady shear
flow as the simplest example.
5.2 CONSTITUTIVE EQUATIONS FOR SIMPLE SHEAR
Let us consider the general set of constitutive equations (1.71H1.76) for
the case of simple shear flows. In accordance with Fig. 4.1 and formulae
(4.1) and (4.2), the kinematic matrices take the form:
-1
0]
[0
0
1
e=~ [1 0 0 ,(J)=~ 1 o ~],
o 00
0 o0

Ck

~],

0 0 1

= [:::::

:::::

~]
(5.1)

Shear Deformations in Elastic Polymeric Liquids

115

where y is the shear rate defined with using the stick condition at the wall;
e and (J) are the strain rate and vorticity tensors; Ck is the Finger measure
of elastic strains in the kth relaxation mode.
The incompressibility conditions, detck = 1 is written as:
(5.2)
The matrices
diagonal form:

Ck

can be brought by orthogonal transformation, qk> to a

({Jk

-sin ({Jk

~in ({Jk

cos ({Jk

COS

qk =

0]

(5.3)

1
where Ck is written in the form that takes into account the incompressibility condition (5.2). Substituting the matrices (5.1) into eqns (1.71H1.76)
with an allowance for the eqn (1.59b) and incompressibility condition
(5.2) yields:
(\.11

+(40d- 1 exp(Yk Wk)(Ck - Ck 1)[Ck -Ck 1 +(Ck + Ck 1) cos 2({Jk] =2YCk.12

Ck.12 + (40 k) -1 exp(Yk Wk)(C~ - Ck 2) sin 2({Jk = YCk,22

(5.4)

where an overdot means d/dt and additional relations are:


2Ck,11 = Ck +Ck 1 +(Ck -Ck 1) cos 2({Jk,
2Ck,22 =Ck +Ck 1 -(Ck -Ck 1) cos 2({Jk

2Ck,12 = (Ck - Ck 1) sin 2({Jk


(5.5)

Here the same notations were used as in Section 1.8. The set of
equations (5.4) and (5.5) is closed relative to two unknown sets of
variables, Ck and ({Jk> if the shear rate y is given.
Due to eqn (1.17), the elastic potential for each Maxwell mode is given by:
(5.6)
where Ji,k is the elasticity modulus and nk is a dimensionless constant.
Then the elastic energy for the simple shear flow of a polymer liquid is:

W=

W k = LJi,k/(2nD(~k +Cknk_2)

(5.7)

and according to eqn (1.96), the stress tensor is generally represented by:
(J

= - pl5 +

I
k

(Ji,k/nk)Ckk + 2s11e

(5.8)

116

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Taking into consideration that C~k = qk . ~k . qk 1, we can easily represent


the components of the stress tensor (5.8) via Ck and <Pk as follows:

1
0"11 = - P + :2 I CUk/nk)[c~k +Ck nk +(C~k -Ck nk ) cos 2<pk]
k
0"22

1
= - P + :2 I CUk/nk)[c~k +Ck nk _(C~k -Ck nk ) cos 2<Pk]

(5.9)

0" 33 = - P + I /lk/nk>
k

1
0" 12 = :2 L (/lk/nk)( C~k - Ck nk) sin 2<pk + S'1Y
k

Equations (5.4), (5.5) and (5.9) form a closed set for the description of
the isothermal shear flows.
As shown in Chapter 4, the following quantities are commonly measured in simple shearing: the first, 0"1> and second, 0"2, normal stress
differences, and the shear stress, 0"12' They are represented as follows:
0" 1 == 0" 11 - 0"22 = I (/lk/nk)(~k - Ck nk ) cos 2<Pk
k

1
0"12 =:2 I(/lk/nk)(c~k -Ck nk ) sin 2<Pk +S'1Y
k

Now let us consider separately the case nk = 1 which is mainly used for
calculations in this chapter and corresponds to the classic potential (1.47) of
the theory of high elasticity. This potential in simple shearing is of the form:

W k = (/lk/2)(Ck.11 +Ck.22 -2)

(5.11)

In this case, there is no need to consider the tensor of elastic strains Ck in


the principal axes and the whole set of equations is simplified.
Let us introduce the dimensionless magnitudes and parameters using
arbitrarily chosen parameters /l (of dimensionality of modulus) and fJ (of
dimensionality of time):

(5.12)
l=yfJ,
Ak =fJk/fJ,
Vk =/lk//l,
Vk = '1k fJ/('1fJk) = '1k/('1Ak))
Substituting the kinematic tensors (5.1) into the constitutive equations
(1.72H1.76) with due regard for eqns (5.2) and (5.12) yields: 65
7: = t/fJ,
(/lk = '1k/fJ k>

22kc\,ll +(C~,ll + C~,12 -1) exp(Yk W k ) = 4Ak f'(7:h, 12


2AkCk,l1C\,12 +Ck,12(C~,l1 +C~,12 + l)exp(}'k Wk)=2Akf'(7:)(1 +C~,12)'
ck,22=(l+C~,u)/Ck,l1
(k=1,2, ... ,N)
(5.13)

Shear Deformations in Elastic Polymeric Liquids

117

where an overdot means d/dr. Equation (5.13) is the set of N couples of


equations for determining the recoverable strains Ck,ij in the kth nonlinear
relaxation mode. If the values of Ck,ij have been found from the set, then,
in accordance with eqns (1.71), (1.72), (5.10) and (5.11), the system of
stresses and elastic potential in the simple shear under study are given by:
O"T =(0"11 -0"22)/f-l=

L Vk(C~,l1 -C~,12 -l)/Ck,l1


k

O"~ =(0"22 -0"33)/f-l=

L Vk(Ck,22 -1)
k

O"T2 =0"12/f-l=
W*

(5.14)

L Vk Ck,12 + s l L VkAk
k

= W/f-l = L Vk(Ck,l1 +Ck,22 -2)


k

As a rule, the influence of exponential factors on the relaxation time is


insignificant, meaning that 11k W k ~ 1 and for a number of cases, the set
(5.13) can be reduced to that with 11k =0. In this chapter, numerical
calculations are made with the equations where nk = 1 and 11k = O. The
calculations with nk =F 1 and Yk =F 0 (see Chapter 7) in the common case of
rheometrical flows where Deborah numbers are not so high show that the
qualitative picture of viscoelastic shear flows does not change substantially. But these effects should be taken into account in high De shearing
and in simple extension, at least for the polymers considered in this
chapter.
5.3 SHEARING IN THE LINEAR REGION OF DEFORMATIONS
5.3.1 Experimental Results
In the region of linear strains, the shear experiments with polymers are
commonly carried out either in the dynamic mode with low-amplitude
shear oscillations, or in the mode of constant strain rate.
As discussed in Section 4.2, the complete strain Y in dynamic experiments is assigned by the real part of the expression:
y=e'exp(iwt)

(5.15)

where t is time, e is the strain amplitude and w is the frequency. For the
cone-plate instrument, shown in Fig. 4.3,
(5.16)

118

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where Ll m and h are the maximum linear displacement and the gap at a
fixed radius r; I/Im and a(l/Im' a ~ 1) are the maximum angle of the plane
twisting and the angle between the cone and plate. Note that the value of
8 does not depend on the radius r.
In the dynamic experiments, the tangential stress CT 12 depending on
time is measured. In the linear region, this stress should be proportional
to 8 (which is checked experimentally) and generally shifted in phase by
an angle cp in relation to strain y(t). In steady oscillations, the phase shift
cp depends only on the frequency wand properties of the liquid. For
instance, in the case of an elastic body, cp = 0, for the Newtonian liquid,
cp=n/2, and for a viscoelastic medium, O<cp<n/2. Using the functions
cp(w) and G(W)='CT12/8 obtained in the experiments, which are independent of amplitude of deformation 8, we can calculate the moduli of
elasticity (storage) G' and losses G" represented in eqns (4.18) and (4.19) as
follows:
G' =Gcoscp,

G"=Gsincp

(5.17)

The moduli G' and G" are called the real and imaginary parts of the
complex dynamic modulus G* = G' + iG". The values of the rheological
parameters and characterisations of polymeric fluids are presented in
Table 5.1. The dependences G'(w) and G"(w) obtained in experiments
with melts and solutions of polydisperse industrial polymers (fluids 1-3 in
Table 5.1) are shown in Fig. 5.1.
When shearing at constant strain rate y within the linear region of
deformations, the ratio CT 12/Y, called the effective viscosity, does not
depend on y. Typical experimental plots CT 12/Y versus time t are shown
for a polyisobutylene melt (see Table 5.1) by the curve of asterisks in
Fig. 5.2. The descending part of the curve of asterisks in Fig. 5.2
corresponds to the relaxation process after cessation of steady-state flow.
Unlike polymer solids, in polymeric liquids under relaxation the stress
drops down to zero with time. In the linear region of viscoelastic
deformation of polymeric liquids, there is a useful relation between the
dependences CT12(t) for start-up flow and relaxation after cessation of
steady-state flow:
(y = constant)

(5.18)

where t is the time counted off from the beginning of shear, r is the time
counted off from the instant to when the relaxation process starts.
Sometimes the process of relaxation is measured after an instantaneously assigned shear strain Yo. The time dependence of the modulus

NN

in polychlorinated
biphenyl Arodor
1248 at 30C)
Polyisobutylene P-20

(M~6.7x105)

Polyisobutylene P-20
at 99C M ~ 105
Y/= 1.1 X 106 Pa.s- I
0=7x10 2 s
Ge = 1.57 x 103 Pa
32.6% solution of
polystyrene

40% solution of butyl


rubber in transformer oil at 22C
(Y/ = 1.9 x 103 Pa.s
0=3.9 s,
Ge=600Pa)
LDPE (at 150C)
p=918kgm- 3 ,
0=5x 104 Pa.s- 1
at 20C
Polystyrene at 170C
(Hoechst N 4000 V)

Material fluid

Table 5.1 Model parameters

Dynamic

Relaxation after
instantaneously
assigned shear
-strain
3

Dynamic

Constant shear
rate

N
of modes

Dynamic

Dynamic

Type of test
in the linear
region of
deformation

2.736 x 103

5.225 X 104
7.735 x 104
9.355 X 102

3.56 X 106
1.30 x 106
1.49 x 106

1.00 X 103
1.00 X 10 1

7.7 X 105
3.3 x 105

8 x 10- 4

6.56 X
3.61 x
1.00 X
5.00 x
2.00 X
1.60 X
2.00 x
2.00 X
1.10 x
8.00 x

2.13 X 104
2.23 x 104
3.62 x 10 3
4.68 x 102
2.02 X 104
4.24 X 104
3.06 X 104
9.26 X 103
7.46 x 102
1.03 X 102

5 x 10- 4

9.91 x 10
2.27 x 10
2.55 X 10- 2

9.02 X 102
1.86 X 102
1.61 X 10 1

10 1
10
10- 1
10- 3
102
10 1
10
10- 1
10- 2
10- 4

8.0x10- 1
1.6 x 10
LOx 10- 1

1.70 x 103
1.03 X 103
1.03 x 103

1.8 x 10- 1
5.88 x 10- 2

Ok
(s)

Y/k
(Pa s)

5.27 X 10 3
3.41 X 104
3.67 X 104

3.95 X 10 3
6.99 X 103
9.25 X 104

7.7 x 102
3.30 X 104

3.24 x 102
6.18 X 103
3.62x 104
9.36 X 104
1.01 X 102
2.66 X 103
1.43 X 104
4.64 x 104
6.78 X 104
1.29 X 105

2.12 X 103
6.40 X 102
1.00 X 104

J.lk= Y/k/Ok
(Pa)

247

209

38

219

274

65

Reference no.

120

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

-f

-II

-3

-2

-f

tfJ,ffU, if

I(.

~'

3
~
.." 2
f

-5

-If

-3

-2

S'

-f

~ If

~. J

~2

0-(

0-2
t; -3

1
-5

-4-

-3

-2

-{

topfU, rf

Fig. 5.1. Plots of dynamic moduli G' and G" versus frequency of shear oscillations w. Symbols 1, 2 and 3 corresponds to fluids 1, 2 and 3 in Table 5.1.

"

If ......

....

...
.

{<

Fig. 5.2. The time dependences of


effective viscosity II dt)/y in linear
region of deformations for start-up
flow and relaxation after cessation of
steady state flow (fluid 4 in Table 5.1).

Gr = (J 12iYo for polystyrene solution (see Table 5.1, fluid 5) is demonstrated in Fig. 5.3. Once again, the relaxation modulus Gr(t) does not
depend on the value of Yo in the linear region of deformations.
5.3.2 Theoretical Considerations

In the linear region of deformation, the dimensional constitutive equations (5.13) and (5.14) take the form of eqns (1.84):
(J12

= L(Jk,12
k

+srry

(5.19)

121

Shear Deformations in Elastic Polymeric Liquids

Fig. 5.3. Time dependence of relaxation modulus G,(t) in the linear region of
deformation for relaxation following an
instantaneous jump of shear strain (fluid
5 in Table 5.1).

20~----~----~2----~~-
(or

t,

where 0"k,12 is the shear stress in the kth Maxwell element and 0"12 is the
total shear stress. In a study of dynamic deformation (5.15), the shear rate
is:

y= iWB exp(iwt)

(5.20)

Then searching for the steady oscillatory solution of the first dynamic equations in eqns (5.19) in the form: 0"k.12 = O"~ exp(iwt), we an easily obtain:
o

O"k

= Bilk 1

G:

iW(Jk
*
. (J =BG k

+IW

where
is the complex dynamic modulus in the kth relaxation mode.
The real and complex parts of
are of the form:

G:

,
(W(Jk)2
Gk = Ilk 1 + (w(Jd 2'

(5.21)

Thus for the whole system,


G'=

IGi.:,
k

G"= IG~+s1'fW

(5.22)

In shearing from the rest state with a constant strain rate


yield:

y, eqns (5.19)

(5.23)

122

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

When starting the stress relaxation after cessation of a steady flow with
shear rate y, the shear stress 0" 12 is instantaneously decreased by the value
S17Y which corresponds to some rapid relaxation processes not taken into
account by this model. Then
0" 12 =

YL 17k exp( - tl(}d

(t>O)

(5.24)

where t is counted off from the beginning of relaxation experiment.


If the instantaneous shear Yo is imposed on a resting viscoelastic liquid,
the relaxation of shear stress occurring after that is described as follows:
Gr(t) =

0" n(t)/yo =

L (17k/(}d exp( -

tl(}k)

(5.25)

5.4 EVALUATION OF RHEOLOGICAL PARAMETERS

The model under study contains a constant s, and N couples of constants:


moduli f.1k and relaxation times (}k; the viscosities 17k being represented as
17k = f.1k (}k' As follows from the previous section, these can be evaluated in
the linear region of deformations. Simple shear tests are usually employed
for this purpose. These are mostly dynamic experiments65.175.207,208 or
experiments with given constant shear rate. 38 In the first case, the
experimentally obtained dependences G'(w) and G"(w) are used. In the
second case, the time dependences of shear stress 0"12 obtained in
experiments either for start-up flow or for relaxation after cessation of
steady-state flow are used. Sometimes, the data on stress relaxation
obtained after step-shear deformation of polymeric liquids are also
used. 209 ,210 Respective analytical dependences are given by eqns (5.21)(5.25). In order to find the parameters f.1k and (}k, the fitting procedure is
used. Hereafter the curves calculated with fitted parameters are shown by
solid lines and experimental results by dots. Some results of the fitting
procedure are demonstrated in Figs 5.1-5.3, while Table 5.1 gives
information about tested polymeric liquids and selected parameters of
model. In Sections 5.5-5.9, we will discuss the experiments with the same
liquids in a nonlinear region of deformation. Also, somewhat different
approaches were used to determine the parameters of the model, but
these do not exert an appreciable influence on the calculations in the
nonlinear region of deformation. In Refs 38 and 65 the data obtained
from dynamic experiments were approximated analytically with a minimal number of Maxwell modes, N. The fit was valid over the largest

Shear Deformations in Elastic Polymeric Liquids

123

range of frequencies w with a predetermined, rough enough, accuracy. It


was found 38 that the region of w where the best fits of dynamic data were
achieved, corresponded to the best descriptions of experimental data in
terms of y for other, nonlinear regimes of deformation, if w is proportional to y. In Ref. 65, the dynamic moduli G'(w) and G"(w) were fitted in
the low-frequency region, w-+O, where, due to eqns (5.21) and (5.22), the
asymptotes
G' ~ w 2

L 11k (}k = w 2 11(),


k

G" ~ w L 11k = W11


k

hold true.
In the case of 40% butyl rubber solution in transformer Oil,65 two sets
of parameters have been selected as related to two fittings of G'(w) and
G"(w) (see Fig. 5.1, fluid 1). The first approximation (N = 1, solid lines)
corresponds to the minimum number of parameters (8=0.18,
11 = 1.7 X 103 Pa s-1, () =0.8s). The second approximation (N =2, dashed
lines) describes the dynamic moduli in a wider region of w then the first
one. It should be noted that the values of (), {l and 11 obtained from the
fitting in the first approximation, were used in eqn (5.13) to introduce the
dimensionless variables and parameters. Furthermore, in this chapter, we
denote the first and second approximate results of calculations, related to
fluid 1, by solid and dashed lines, respectively. Also, a rather rough
description of linear start-up flow under a constant shear rate was used in
Ref. 38 to approximate the data in Fig. 5.2. In Refs 207 and 210 the
parameters {lk and (}k were chosen so that the theoretical dependencies
might describe the data with an accuracy of the experiment over the
whole range of existing frequencies and times (see fluids 2 and 3 in Fig.
5.1, and Fig. 5.3). As shown in Table 5.1, this requirement leads to an
appreciable increase in the number of rheological parameters involved.
As a rule, the rejected relaxation mechanisms with small relaxation times
were replaced by a single viscous term with parameter s#O (see Table 5.1,
fluids 1-3).
When evaluating the parameters (}k and {lk, a relaxation mechanism
with a large relaxation time (}1 and a small viscosity 111 as compared to
total viscosity 11, can sometimes be lost. This is the so-called 'large-scale
relaxation mechanism with a small weight' that may play an important
role in nonlinear relaxation phenomena (see Chapter 7). In order to
reveal such a relaxation mechanism, dynamic tests with very small
frequencies ((}lW~ 1) or normal stress measurements at very small shear
rates ((}ly~l) should be carried out, and these measurements may be

124

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

quite complicated. In simple shear, the effect of the relaxation mechanism


is usually insignificant, but it could be important in extension.
Now let us briefly discuss the parameters defined in the nonlinear
region of deformation. As mentioned previously, in the medium range of
nonlinear shear behaviour, we can neglect the effects of recovery strain on
the relaxation times and assume that the parameters f3 and y in eqns (1.58)
and (1.59b) are negligible. These, and particularly y, may be important in
the extension; however, for the large-scale relaxation mechanism for
which y~1O-5cm3 mole-i. To satisfy the qualitative description of
normal and shear stresses, the constraints, 1 < nk < 2, are imposed on the
values of parameters nk in power-like elastic potential (5.6). If the
large-scale relaxation mechanism is not taken into account, the constants
nk can be considered to be independent of k. In the moderate region of
shear deformation, where the neo-Hookean approximation of stresses is
valid, parameter n = 1, and we will use this value in calculations made in
this chapter. If more nonlinear effects should be considered, it is enough
to increase the value of n slightly. For instance, in order to describe a
substantial increase in elongational viscosity it is taken into account in
Chapter 6 that n ~ 1.1. This small increase in n affects the shear calculations little when compared to n= 1. Usually the constants nk are convenient to select by the dependence of the first normal stress differences on
shear rate when the constants ek and J1k have been found.
5.5 HOMOGENEOUS SHEARING UNDER A GIVEN SHEAR
RATE (START-UP FLOW AND STRESS RELAXATION)
5.5.1 Experimental Results

Shearing tests under a given constant shear rate y can be achieved in a


parallel plate device (Fig. 4.1) by assigning a constant velocity v of a
mobile plate, and in rotation devices (Figs 4.2 and 4.3) a constant angular
velocity Q of a mobile unit (cone and cylinder).
Figure 5.4 demonstrates typical dependences (fluid 1, Table 5.1)65 of
shear stress (T 12 on dimensionless time L = tie in start-up flows with
various dimensionless constant shear rates r = ye. Hereafter, the data are
represented by points. The remarkable feature of these data is that in
nonlinear shearing, the curves pass through a maximum before reaching
the steady-state value corresponding to steady shear flow. Sometimes in
experiments with higher shear rates y, a small minimum is observed
following the maximum, and only after that is steady flow reached. The

125

Shear Deformations in Elastic Polymeric Liquids

a)

Gt.6fi.... -,
~I

'I

l
I

................. _

B)

------!

6
C)

'lQ+ f

'l'

'&'0+ 2

Fig. 5.4. Theoretical predictions (curves) in comparison with experimental data


(points) for the dependences of shear stress (1), first and second normal stress
differences (2 and 3, respectively) on dimensionless time 1: in start-up shear flow
with a constant shear and subsequent relaxation. The values of dimensionless
shear rate r are (a) 2.88, (b) 7.25 and (c) 9.13. 65

126

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

higher the shear rate y, the more intensive is the maximum in 0"12 and the
earlier it is achieved in time (and later in the deformation, yt), as well as
the steady flow (see, e.g. Refs 1 and 65). At small values of y the maximum
on the dependences 0"12(t) degenerates as it goes to t-+oo and the curves
begin to increase monotonously (Figs. 5.2).
Usually, if the second start-up flow follows the cessation of steady-state
flow after the duration of a relaxation process shorter than a characteristic relaxation time e (which is approximately equal to the time to reach
the steady flow), the maximum 0"12 value or the repeated flow may be
somewhat smaller than the initial one. For some liquids, usually colloid
systems and polymers filled with active filler, the delay or 'rest' time
needed to achieve the initial value of maximum 0"12 in repeated shearing,
substantially exceeds the time necessary to reach the steady flow (see, e.g.
Refs 211-213). This phenomenon, known as thixotropy, is usually associated with the time evolution of such structures in these liquids as the
secondary network created by interactive (usually attractive) colloidal
particles. 214 Such liquids are not generally considered below.
Figure 5.4 also shows the dependences 0"12 (-r) after cessation of steadystate flow at instant ro.65 This is the process of stress relaxation that is
related to intramolecular motion in the polymeric liquids after stopping
the macroscopic flow by the fixation of a mobile plate. The higher the
shear rate in steady flow, the higher is the speed of the relaxation
processes. 215 ,216 Only the steady-state values were obtained for the
second normal stress difference. It should be reminded, that along with
the tangential stress in simple shearing of polymer liquids there also exist
normal stresses. The respective data of the first normal stress difference
O"l(r), defined by eqn (4.4), have qualitatively the same form; steady-state
values and relaxation occurring more slowly72,217,218 than for 0"12(r) (see
Fig. 5.5).173,219 This difference in time behaviour is often as high as
several dozens (see e.g. Refs 217 and 220). As discussed in Section 4.2, the
latter may also be due to improper operation of measuring devices. 172,174
Also, the relaxation after given shear step variations implemented by
different methods finds wide application in shear rheology. Given below are
two examples that are of interest owing to their subsequent comparison
with the calculations. Relaxation dependences of the effective shear
modulus Gr (r)=0"12(r)/yo after given shear steps173,274 demonstrated in
Fig. 5.6 show that the higher the value Yo, the lower is the respective
dependence. Using the optical method, the relaxation of shear and normal
stresses were measured after two subsequent shear steps Yo and Y1 imposed,
with the time interval to between them.209 These data are shown in Fig. 5.7.

127

Shear Deformations in Elastic Polymeric Liquids

~
4

....

~""
~

0-1

~3

0-2

<i;} 2

2
f

-2

-1

-I

(Wi, ;/

Fig. 5.5. Time dependences of shear


stress 0" 12(t) (curve 1) and first normal
stress difference 0"1(t) (curve 2) in the
relaxation of fluid 2 (Table 5.1) after
cessation of steady flow with shear rate
y= 1. 173 ,274

~
0

t, ;/

Fig. 5.6. Time dependence of relaxation modulus Gr(t) in relaxation following an instantaneous jump Yo of
shear strain (fluid 2 in Table 5.1). 0,
Yo=1.9; D, yo=9.5; 6, yo=30.9. 173 ,274

5.0

~'t;/
I
~

4.0

~
~~ 3.0

"
2.0
0

1.0

2.0

t0tl

t,

J.O

4.0

;f

Fig. 5.7. Time dependences 209 of shear stress 0"12(t) and normal stress 0"11 -0"33
in relaxation after two consequent strain jumps with Y01 = 1.47 and Y02 = 1.43 with
time interval between them to=21Os (fluid 5 in Table 5.1).

The most common, however, are the dependences of shear stress 0'12(Y),
and the first O'l(Y) and second 0'2(Y) normal stress differences obtained at
steady-state flow. These are shown in Figs 5.8 65 and 5.9. 38 ,159 Comparable results were obtained in Refs 1 and 221. The plot 0' 12(Y) is usually
called a 'flow curve'; the ratio if = 0' 12 (Y)/Y being the effective shear
viscosity. For polymer melts and concentrated solutions, the effective
viscosity if decreases with growing y and reaches the constant maximum

128

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

fO

15;/2,6;;,-02*'

G", (C*): (G")'

'/

'"'"

'"

'"

0- f - 6,1 1 p=l5l~

.-2 - (91 1~ =4;/2


A-3-6'/~ -(6*)
o -1j-VG,z+G'/~=G*
0-5- G' 1 fI= (G*)'

f'

102~----------~------------~--------'-~~*~
10-(

10

10 2

Fig. 5.S. The plots of flow curve, the first and second normal stress differences
versus dimensionless shear rate r (curves 1, 2 and 3, respectively). Theoretical
predictions are shown by lines, and data for fluid 1 (Table 5.1) by dots. Data for
G' and G" versus dimensionless frequency of oscillations w* = Ow, are shown by
dots 3, 4 and 5. 65

value 11 equal to the Newtonian viscosity only at very small shear rates y.
For the commonly studied polymer melts, 11 ~:::::104 -10 7 Pa s -1,
The first normal stress difference is positive while the second is
negative. When y~O the shear stress and both normal stress differences
have asymptotes (see, e.g. Ref. I);
i= 1,2

The constants ei are known as the initial normal stress coefficients. These
asymptotes are seen in Fig. 5.8. The dependences O'u(y) and O'l(Y) increase
monotonously. At higher values of y, when they usually increase more
slowly than a linear function of y, 0'1 (y) > 0' 12(y).l There is also a lot
evidence in the literature that 10'2(y)1 also monotonously increases with
growing y and the ratio 10'2(Y)I/0'1(y);::::0.1--OA. 161 ,222-230
Experiments show that at a low shear rate y, the flow curves 0' 12(Y)
coincide with the loss modulus G"(w) while one-half of the first normal

129

Shear Deformations in Elastic Polymeric Liquids

2.0

1.0

0.5

1.0

Fig. 5.9. Experimental plots of shear stress 0"12 (0), the first 0"1 ( L:.) and second
0"2 (0) normal stress differences versus shear rate y: (a) fluid 4 (Table 5.1),38 (b)
32.5% PS solution in Arochlor 1232. 159

stress difference, 0" 1(y)/2, coincides with the storage modulus G'(w) (see
Fig. 5.8). These are known as the Cox and Merz rules. 231 In a wider range
of y and w, the relation commonly fulfilled is:
0" 12 (y) =

IG*(w) Iy=ro

5.5.2 Theoretical Considerations


Let us study the transient shear flow from the rest state ('t"=0) to
the steady-state flow ('t"-+oo), which is also called the 'start-up flow'.

130

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The initial conditions for the components of Ck tensors are as follows:


(5.26)

In this case, set (5.13) with initial conditions (5.26) is easily studied on the
phase plane {Ck,l1; Ck,12}' The only steady point on the phase diagram is a
stable focus which means that the solution reaches the steady flow regime
with decaying oscillations; Ck, 12 Cr) and Ck,l1 (r) passing through pronounced maxima when r = By is high enough. With the increase in r, these
maxima become more intense and shift to the region of small r. The same
holds true for 0'12(r) and O'l(r). This confirms that the set can describe, at
least qualitatively, the experimental data discussed in the previous section.
When r-> 00, the solution approaches steady state which could describe
the steady shear flow of viscoelastic liquids. With Yk = 0, the stationary
solution of the set (5.13) takes the form:
Ck,12 (00) = 2A kr /(1 + Xk),

Ck,11( 00) = Xk [2/(1 + Xk)J1/ 2,

- O'!( 00) =

L Vk {1- [2/(1 + Xk)] 1/2},


k

W*( (0) =

L vk[(2 + 2Xk)1/2 -

2]

(r=By,

xk=(1+4A~r2)1/2)

(5.27)

where 0'1, a! and 0'12 are defined by eqns (5.14). In the general case, the
set (5.13) cannot be solved analytically even when r=constant. The
results of its numerical calculations are given below in comparison with
the data.
Unlike the start-up flow, there is an analytical solution of eqns (5.13)
when Yk = 0, which describes the process of stress relaxation after cessation of flow at instant roo Studied here is the stress decay at r > ro with
r=o due to eqns (5.13) with the initial conditions:
(5.28)
The solution of the set (5.13) is easily found and then the normal and
shear stresses can be expressed as follows:

* _"

0'1 -4

VkPkl/!k exp( - ed
(1- I/!~ exp( - 2ek))(p~ +4)1/2

131

Shear Deformations in Elastic Polymeric Liquids

(5.29)

Formulae (5.29) describe in particular, the process of relaxation after


cessation of the steady-state flow when the values c~,ij = Ck,ij (<Xl) and the
latter are given by eqns (5.27).
The expressions for the stresses (5.29) in the process of relaxation
possess the following properties:
(i) at the starting instant of relaxation, r=ro+O, the shear stress 0"12
has a step-wise decrease by the value srL k VkAk which models very
fast processes of shear stress decay, so the shear stress attenuates
faster that the normal ones;
(ii) at the starting instant of relaxation, r = ro + 0, the rate of stress
relaxation is at a maximum;
(iii) a linear 'relaxation spectrum' is associated with every nonlinear
kth relaxation mode, which is rapidly attenuated in the highfrequency region of relaxation as follows from the expansion:
t/J exp( - ~k)
1-;2 e p(-2~k) =
k x

L t/Jiexp(-~d)
j~l,3,

This 'spectrum' is effectively truncated on the low-frequency side.


All these features are in good qualitative agreement with experimental
observations.
Let us also consider the case when the shear strain is assigned by a
step-wise variation 207 ,21o with value Yo at an instant to. In this case,
y=yo<5(t-t o ) where <5(t) is the Dirac b-function. To describe these
experiments, parameter s in eqns (5.14) should vanish. As shown in Ref. 207,
the general shear step-wise variations of Ck tensor at instant to is given by:
D+YoB
B

~]

(5.30)

132

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where A, Band D are the initial conditions for Ck.ij' In particular, when
the step-wise variation is imposed on the rest state, A = B = 1 and D = 0,
and eqn (5.30) is simplified to the perfect elastic response:

Yo 0]
1

In this case, the relaxation process in simple shear


following:

(5.31)

IS

described as

k)
-----=2:-'--=--'----'----'-k ))

(5.32)

O'12(t) = Lllkak(t),
k

ak (t ) =

4yo exp( - t/fJ


4+yo(1-exp( -t/fJ

The second relationship in eqns (5.32) was noticed in Ref. 232. Also,
two-step shear deformations were considered in Ref. 207 in comparison
with literature data.
5.5.3 Discussion of Results
In Figs 5.4-5.9 the calculations are represented by lines, and the data, by
symbols. For fluid 1 (see Table 5.1), a three-parametric version is shown
by solid lines and the five-parametric one, by dashed lines.
The comparison between theoretical predictions and data demonstrates generally their quantitative agreement. To describe in more detail
the experiments at low time and high shear rates, all the smaller
relaxation time modes have to be retained.207.21o
Figure 5.10 also shows the three-parametric theoretical predictions
of dimensionless elastic energy W* for fluid 1 (Table 5.1) in start-up
shearing with a constant shear rate and in the subsequent relaxation
process. These dependences are similar to those for the stresses shown
in Fig. 5.4. The behaviour of dissipative function D(t) for this model
in simple shear can be also calculated according to the general expressions (1.43) and (1.44), and its behaviour is analogous to that of
W(t).

Additionally, the rheo-optical measurements known from the literature were compared in Ref. 210 with the predictions of the present
theory. Namely, the birefringence An and the extinction angle X for
stress growth in start-up flow with subsequent relaxation were cal-

133

Shear Deformations in Elastic Polymeric Liquids

2
./

./

-'-'/'-'-if1
.

2
Fig. 5.10. Theoretical predictions of time dependences of elastic energy W*(r) in
start-up flow with subsequent relaxation, r=2.88, 7.25 and 9.13 (curves 1,2 and
3, respectively).

_-----'\r--

120

100

00

1=0-01551

20

40

80

120

Time (5)

(al

0
0

Y=0-0155-1
002
0-033
0'05
007

0>

x 30

25~--~-----L----~----~--~-----L~~--

100

200
Time (5)

300

(b)

Fig. 5.11. (a) Birefringence ~n and (b) extinction angle X versus time following a
sudden imposition of shear flow at different shear rates. 210 Solid and dashed
curve correspond to 6- and 4-mode fit to theory; points represent the data of
Gortemaker et al. 219 for fluid 3 (Table 5.1).

134

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

100~--~--~4~O~--~--8~O~--~--~12~O~~~O~---L---J40

Time (5)

Fig. 5.12. Extinction angle X versus time following a sudden imposition of shear
flow (upper curves) together with relaxation after the attainment of steady-state
flow (right curve) or after a flow duration of 20,40 or 80 (other curves).21 0 Curves
represent theory and points represent experimental data by Gortemaker et al. 219
for fluid 3 in Table 5.1.

culated according to the model and common formulae Lln= C[40"I2 +


(0" 11 - 0"22)2r/2, X= tan -1[20" 12/(0" 11 - 0"22)] and compared with previous
experiments. 219 Here C is the stress-optical constant. For fluid 3 (see
Table 5.1), where C =4.2 x 10- 9, the results of the comparison are shown
in Figs 5.11 and 5.12. Other comparisons, including rheo-optical studies
and a comparison with predictions of other theories, can be found in Refs
207 and 210.
5.6 PARALLEL SUPERPOSITION OF OSCILLATIONS ON
STEADY SHEAR FLOW

5.6.1 Experimental results


The superposition of shear oscillations with frequency 0) on steady shear
flow with shear rate y have been extensively studied experimentally and
theoretically.65,208,233-255 Though the most common is the one-dimensional scheme of experiment known as parallel superposition, a twodimensional scheme or orthogonal superposition is also used. 237 ,24o The
latter will be considered separately in Section 5.7. The most complete
experimental study of polymer solutions in parallel superposition is
described in Ref. 238.
The most convenient instrument for parallel superposition experiments
is the cone-and-plate rheometer sketched in Fig. 4.3 with the angular

135

Shear Deformations in Elastic Polymeric Liquids

velocity of rotated cone given by:


n(t) =

no + BW cos wt

where no, Band ware constant. The total strain rate y(t) is represented as
n(t)/O(; its constant component is Yo = nolO( and the strain amplitude of
oscillations is e = BIO(. The torque on the disk from which the shear stress
is calculated, is recorded in time simultaneously with the oscillating strain
component Y1 =e sinwt. Constant and variable components of shear
stress 0"12 were measured in these experiments, and then the phase shift
({J12 between oscillating stress and strain components as well as the
dependence of complex dynamic modulus Gt2 on wand Yo was also
determined. In these experiments, the variable component of the shear
stress was linear in e. The constant component of the thrust, and
therefore, of the first normal stress difference 0" 1> was also measured but
due to the defects in the compensating scheme mentioned in Section
4.2.5,65.172 it was impossible to detect reliably the variable component of
0"1

Figures 5.13 and 5.14 demonstrate the amplitude ratios between


variable components of shear stress and strain, G12 , and the phase angle

0---

--o{)---

-O---r::r-_:J

"-

---0---0.:: ....

0-1
A - 2

90'

~"-

10 2

--

~2

CfE , Pa

"-

0-3
10

~""-

"

to.

"1- 1

Fig. 5.13. Dynamic modulus G 12


versus steady shear rate Yo at various
frequencies w in parallel superposition
of small amplitude oscillations on
steady shear flow. Lines represent theory and dots represent experimental
data for fluid 1 in Table 5.1: w=0.397,
0.993 and 6.28 s -1 (curves 1, 2 and 3,
respectively.

50

fQ-f

A-2
0-3

0 0

9'

ry/

.- '" 0-1

.... 0

"-

9'

700
~

,,-

to,;ff

to

Fig. 5.14. Phase shift CfJ12 versus


steady shear rate Yo at various frequencies w in parallel superposition of small
amplitude oscillations on steady shear
flow. Lines represent theory and dots
represent experimental data for fluid 1
in Table 5.1: w=0.397, 0.993 and
6.28 s -1 (curves 1, 2 and 3, respectively.

136

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

between them, ({J12, depending on the Yo at fixed w. Here the experimental


data are denoted by symbols and theoretical predictions, by lines; solid
and dashed lines designating the first and second approximations, respectively. At small values of Yo, the values of Y12 and ({J12 correspond to
those obtained in pure oscillatory experiments (see Section 5.3). The plot
({J12 versus Yo has a maximum whose value exceeds 90. With a further
increase in Yo, the angle decreasing, tends asymptotically to 90. Note
that in pure dynamic experiments when Yo=O, 0<({J12 <90. Figures
5.15 and 5.16 show the plots G12 and ({J12 versus w at a fixed value of Yo.
At large values of w, the dependences G 12 (w) and ({J12(W) tend to the
values corresponding to those for Yo =0. The smaller Yo is, the lower
values of w required to reveal the nonlinear properties of elastic liquid
are. Beginning from a certain value of Yo, the dependence ({J12(W) may
have a maximum greater than 90.
90
f04

Giz , Pa,
70
{oj

50"
tOw,

,,-I

Fig. 5.15. Dynamic modulus G12


versus frequency OJ at various shear
rates Yo in parallel superposition of
small amplitude oscillations on steady
shear flow (compare with Fig. 5.13).
Lines represent theory and dots represent experimental data for fluid 1 in
Table 5.1: Yo = 1.44 and 0.36 S-l
(curves 1 and 2, respectively).

Iii

10

41, ,,-1

Fig. 5.16. Phase shift ({J12 versus frequency OJ at various shear rates Yo in
parallel superposition of small-amplitude oscillations on steady shear flow
(compare with Fig. 5.13). Lines represent theory and dots represent experimental data for fluid 1 in Table 5.1:
Yo = 1.44 and 0.36 s -1 (curves 1 and 2,
respectively).

The values ({J12 and G 12 measured in these experiments are related to


the real and imaginary parts of the complex dynamic modulus by the
common relations:

Shear Deformations in Elastic Polymeric Liquids

137

The plots 1'/'12 = G12 /w and G'12/W2 depending on w at fixed values of

Yo, are presented in Ref. 238. At small w, the plot 1'/'12 versus w approaches
a constant depending on Yo. When w increases, the plot passes first

through a minimum and then through a maximum, approaching further


the dependence of linear dynamic viscosity, 1'/'(w). The higher the value of
Yo is, the lower the path of the respective dependence is. At fixed Yo, the
plots G'12/W2 versus w begin from negative values (first observed in Ref.
236) and then tend to the linear dynamic curve G'(w)/w 2 ; the latter always
being positive, decreasing and tending to zero when w is increasing. In the
region when Gi2 is negative, <P 12 > 90. The following empirical relation
was observed: <P12';:::;1C/2 when w,;:::;yo/2. 236 The data 65 ,208 presented in
Figs 5.13-5.16 are in agreement with the results obtained in Refs 233-236
and 238.
Small constant contributions to the shear stress, 0"12,0, and first normal
stress difference, 0"1,0 were also found in Ref. 208. These appear due to the
interaction between steady shear flow and oscillations. The component
0"12,0 always had negative values and was an order of magnitude lower.
than the shear stress corresponding to the steady-state flow. The normalized plots O"tO=O"l,O//1 versus ro=Yoil (with /1=5 x 102 Pa and
iI = 0.8 s) at fixed values w* = w ii, are shown in Fig. 5.17. At small values
of r 0, the plots are independent of r 0 but with the increase in r 0, all the
plots, independently of w*, approach a common envelope.
Nonlinear shear oscillations without constant shearing component
(r 0 =0) are described in Refs 172 and 239-244. For polymeric melts and
solutions, it was found that while the shear stress depends linearly on the
amplitude 8, the normal stress, 0"1 is proportional to 8 2 have both the
constant and variable components; the latter oscillating with the frequency 2w. With a further increase in 8, the constant component of 0"1
grows more slowly than 8 2 As shown in Refs 172 and 245, the variable
component of 0" 1 is often difficult to measure reliably. Also, a decrease in
1'/' = G12 /w and G'12 with 8 growing was observed in Ref. 248 when
nonlinear shear oscillations were imposed on polymer melts and filled
polymers.
5.6.2 Theoretical Considerations

Theoretical treatment was covered in Refs 65, 208 and 247 and is
therefore a little generalized in this section. We shall use for the theoretical calculations, eqns (5.13) and (5.14) where

r = r 0 + 8W coswr

(5.33)

138

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

A-I

0-4

Fig. 5.17. The constant component aio of the first normal stress difference
versus dimensionless steady shear rate r 0 at various dimensionless frequencies w*
in parallel superposition of small-amplitude oscillations with (8 = 0.5) on steady
shear flow. Lines represent theory and dots represent experimental data for fluid 1
in Table 5.1: w*=O, 0.1, 0.316,1.0 and 10 (curves 1-5, respectively).

Here the dimensionless frequency w is used, scaled by a characteristic


relaxation time e. Because in the constitutive equations under study the
relaxation modes do not interact and Ck,33 == 1, we omit for a while a
lower (mode's) index k in eqns (5.13) and (5.14) and introduce the
notation for the tensor C as a 2 x 2 matrix with the components:
Cll

=1,

C12

=g,

(5.34)

The general structure of the solution of eqns (5.13), corresponding to the


steady oscillations, can be represented in the form of the Fourier series:
00

c=cs+co+

(c m sin mT+Cm cos mT)

(5.35)

m=l

Here Cs is the steady-state solution without oscillations (e~O) given by the


expressions in eqn (5.27) with allowance for the new notations (5.34), and
Co is a constant matrix which appears to be due to the interactions

Shear Deformations in Elastic Polymeric Liquids

139

between oscillations and steady flow, as well as between the oscillations


themselves. The constant matrices Cm in eqn (5.35) depend on r 0 and the
parameters of oscillations e and w.
We assume initially that the amplitude of oscillations e is small, and
consider the results and numerical calculations for arbitrary e at the end
of this section. When e ~ 1, we can further expand the Fourier coefficients
Co, Cm and cm in the asymptotic power series of e:
00

{CO'Cm'Cm }=

L {Con,Cmn,Cmn} en

(5.36)

n=1

where now the matrices Cmn depend on the constant shear rate r 0 and the
frequency w. The same structure shown by eqns (5.35) and (5.36), holds
for the components f, g, and h of matrix C and the system of stresses
represented by eqn (5.14). By passing from the single to the multimodal
case we use the obvious transformations:
(5.37)

Upon substituting the expansions (5.35) and (5.36), we can always obtain
after tedious but straightforward calculations, the explicit expressions for
the first few terms in the e-expansion.
Consider first, the case when r 0 = O. Then the components of matrix C
are represented as follows:
f = 1 + e2(f 02 + f22 sin 2wr + 122 cos 2wr) + 0(e4 )
g=e(gl1 sinwr+g11 coswr)+0(e 3 )
h= 1 +e 2(h02 +h22 sin 2wr+h22 cos 2wr)+ 0(e 4 )

(5.38)

Thus when r 0 = 0, the normal components of elastic strain have the


expansions only with even powers of e, while the shear component has the
expansions only with the odd ones. The first Fourier coefficients in eqn (5.38)
are calculated straightforwardly and are represented as follows:208.247
w 3 (4+7w 2)
f22 = 2(1 + W 2)2(1 + 4w 2)'
~

~
w 2(8w 4 + w 2 - 3)
f22 = 4(1 + W 2)2(1 + 4w 2)

gl1= 1+w 2 '


h02 =t(9I1 +gI1- 2f02),

h 22 =gl1gl1- f22'

h22 =tWI1-9I1- 2f22)


(5.39)

140

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In this case, the set of stresses (5.14) takes the form:


at = 62 L Vk[fk,02 - hk,02 +(fk,22 -hk,22) sin 2Akw*1'
k

(5.40)
a! = 62

Lvk(hk,02 + hk,22 sin 2llkw*1' + hk,22 cos 2llkw*1') + 0(6

4)

at2 = 6

LVk[gk,ll sin Akw*1' + (gk,ll + AkSW) cOsAkw*1') + 0(6

3)

As seen from eqn (5.39), the expression for at2 in this approximation,
coincides with that known from linear theory of viscoelasticity.
The formulae (5.40) have been obtained in Ref. 72 by using the general
model of 'simple liquid' (see Section 2.2.3) whose particular case is now
under consideration. With the accuracy of up to 0(6 4 ), the constant (a?)
and complex dynamic modulus (Gt) for variable components of the first
normal stress difference is represented as follows:
a? = 6 2G'12(w)

Gt2 =6 2 { G'n(w)-!G'12(2w) + i[G'12(2w)-!G"(2w)]}

(5.41)

Equations (5.41) have been tested experimentally many times (see, e.g.
Ref. 172 and display good agreement with the data.
Also the dependence for the 0'6 can be calculated including the term
0(6 4 ), but the expression is very awkward to demonstrate here. This
dependence, plotted as a? versus 62, shows the linear behaviour when 62 is
small and then the rate of its growth slows down. Such behaviour is in
qualitative agreement with the experiment described in Ref. 172.
Consider now the case of superposition of small-amplitude oscillations
on the steady shear flow (r 0 #0). With the accuracy of 0(6 2 ), the structure
of the matrix C has the general form:
C=Co+6(Cll sin w1'+cll cos w1')+6 2(c02 +C22 sin 2w1' + C22 cos 2w1') + ...
(5.42)
Here the components of matrix Co are represented by eqns (5.27) and
other matrices can be found after straightforward but tedious calculations. The components of matrices Cl and Cl which contribute in the
dynamic moduli are of the form:

_ 4row(l+w2)
11-

K(l+x)

f~ _ 2xr ow(l +W2)


11-

K(l +x)

141

Shear Deformations in Elastic Polymeric Liquids

gl1
h

.J2w 2(W 2 + 2 - x)
1((1 +x)
,

A
gl1 =

2xrow(1 +W2)
1((1 +x)

Ii _ 2goB11 -hOJ11

_ 2g0g11 -hof11
11fo
'

11-

(5.43)

fo

where
(5.44)
Also, the difference between the f02 and h02 components along with g02
which contribute in the measured steady stresses due to oscillations, are
represented as follows:
_ (x 3 -6x 2 -x-2)(gIl +II)
3/2
f02- h024.J2x2(1+x)I/2
+2w(x-1) (f11g11+f11g11)
A

4.J2r 0(2x - x 2 + 1)
1
+x

- (3 + x)(1 + X)1/2(fI 1 +
ro(x-1)

A2

4 (

w x+

1)3/2 A

gl1

nd

g02= 2x(x+1) (gl1 +gl1)+ 2x(f11g11- f11g11)


f11g11 + J11B11
.J2(1 + X)I/2

(5.45)

Similar to eqns (5.40), the formulae for the dimensionless stresses are
represented in the form:

o-t = o-tW0) + e( G'lI sin w*r + G'll cos w*r)


+ e2(o-tq + G1q sin 2w*r + G1q cos 2w*r) + O(e 3 )
0-~=0-~(rO)+e(G21 sinw*r+G 21 cosw*r)
+ e2(0-~q + G2q sin 2w*r + G2q cos 2w*r) + 0(e 3 )
o-t2 = o-t2 (r 0) + e(G12l sin w*r + G'121 cos w*r)
+ e2(o-t2q + G12q sin 2w*r+ G12q cos 2w*r) + 0(e 3 )

(5.46)

Here, G' and G" are the real and imaginary parts of dimensionless
complex dynamic moduli, scaled by elastic modulus ji, for every component of the stress system denoted by lower indices 1,2 or 12; the third
lower index stands for the linear (I) or quadratic (q) term in the

142

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

a-expansion. Also, the first terms in brackets multiplied by a2 stand for the
non-variable addends which appear due to interactions between the
modes of oscillations and steady shearing and between the modes of
oscillations themselves. Though all these characteristics depending on w*
and r 0 are expressed explicitly, 208,24 7 the respective formulae are so
bulky that only the dependences G'12b G'{21 and O"t q are shown here,
which are compared with experiments below. Using eqns (5.42)-(5.45)
and their transformations (5.37) to the multimode case, we can represent
these formulae as follows:
G'12(W*, ro)=

L Vkgk,lb

G'{2(W*,r 0)=

G'l(W*, r 0)=
O"tq(W*, r 0)=

L Vk(fk,11 -hk,l1),
k

L Vk(fk,02 -

L Vdgk, 11 +SAk W *)

G~(W*, r 0)=

L Vk(]k,ll - h

kll )

(5.47)

hk ,02)

To simplify the notations, the lower index'!, was omitted in expressions


(5.47) for the shear and first normal stress difference components of

dynamic moduli.
For the linear dynamic shear moduli G~2 and G~2' the comparison
between the data 208 and theoretical predictions given by eqns (5.43) and
(5.47) with allowance for eqns (5.37), is demonstrated in Figs 5.13-5.16.
The theoretical and experimental data are represented there in the form
of plots qJ12 and G 12 versus w*=wO and ro=YO, where
qJ12

= tan -1(G'{2/G'12),

G12 = [(G'12)2 + (G'{2)2r/ 2

(5.48)

The theoretical predictions are shown in these figures by lines; the first
approximation by solid lines and the second approximation by dashed
lines (see Table 5.1). In Fig. 5.16, the theoretical dependence qJ12 on w*
for the first approximation is shown only in such a region of w* when this
approximation is valid. The theoretical predictions of the data in the
region of higher w* and r 0 values are improved if more relaxation modes
with lower relaxation times Ok are taken into consideration. This was
demonstrated in the calculations of Ref. 247 where three to five relaxation
modes were used to compare the theoretical predictions with many data
available in literature.
The results of calculations 208 ,247 also showed that the amplitude of the
second harmonic for shear stress is smaller by more than an order of
magnitude than the amplitude of the first harmonic. In experiments,208
the second harmonic of shear stress has not even been revealed in the

143

Shear Deformations in Elastic Polymeric Liquids

studied range of w*, r o. It should also be noted that the calculations of


the addend 8 2 0"12q made in Ref. 208 for the first approximation (N = 1)
showed that this quantity is negative and at fixed r 0, has a minimum at
certain value w!(r 0)' With w* = 1 and r 0 = 0.3, the absolute value of this
component is equal to 0.1 G12
Theoretical predictions (N = 1) in comparison with data for the constant component of the first normal stress difference

are presented in Fig. 5.17 208 as plots of O"fo versus r 0 at fixed values of w*
and 8=0.5.
It should be noted once again that the variable component of first
normal stress difference 0"1 is very difficult to measure reliably. Perhaps
this is the reason why the comparisons between the theoretical predictions of present and other constitutive equations (see, e.g. Ref. 247)
revealed a very significant disagreement with the data of Ref. 248.
6
Polyisobutylene

""'<11
III
CL

F
0)

.Q

~3~----_--_L2--------_1~------~O--------L-------~2------~3

Fig. 5.18. Flow curves obtained in oscillatory flow superimposed upon Couette
flow at various strain amplitudes and constant frequency (25 Hz) for material
6 in Table 5.1 at 60C (Fig. 18 in Ref. 247). Curves are for theory; symbols
are experimental data from Ref. 248. 0, 1'0=0.0; X, 1'0=1.0; ~, 1'0=2.0;
,1'0=3.2.

144

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In addition to the analytical solutions for the small-amplitude oscillations imposed on the steady shear flow, the straight numerical calculations of eqns (5.13) and (5.14) with 'Yk=O were also performed for the case
when the large-amplitude oscillations are superimposed on steady
flow. 247 The problem was solved in this work under certain initial
conditions until the steady-state oscillations appeared. Though a lot of
results of the numerical analysis can be found in Ref. 247, the only
comparison with data was made in Ref. 248 because there are few data
available on this subject in the literature. Figure 5.18 demonstrates
the effect of large-amplitude oscillations superimposed on the steady
Couette flow for fluid 6 in Table 5.1. It is seen that the increase in the
amplitude of oscillations results in a significant decrease of the averaged
shear viscosity.
The other theoretical treatments of parallel superposition of oscillations on steady shear flow of viscoelastic liquids can be found in Refs
233-236, 238, 243, 245 and 249-256.

5.7 ORTHOGONAL SUPERPOSITION OF OSCILLATION ON


STEADY SHEAR FLOW

5.7.1 Experimental Results


In this case, an elastic liquid is steadily sheared with a constant shear rate
Yo in direction 1 (Fig. 4.1) while the oscillations with frequency 0) and
amplitude 8 are imposed in orthogonal direction 3.
Until now, this flow was experimentally studied (with polyisobutylene
solutions in cetane) only in Ref. 256, perhaps because it is very difficult to
obtain reliable experimental results. Only the small-amplitude oscillations (8~ 1) were used in these experiments where the dependences of
dynamic storage modulus G23 and dynamic viscosity 1123 = G23 /0) in
orthogonal shearing, on 0) and Yo were studied. The data are shown by
symbols in Figs 5.19-5.21 and demonstrate that the effects obtained,
though less pronounced, were similar to those observed in the parallel
superposition scheme. Also, no negative values of G23 were observed in
the region of small 0), as compared with G'12 for the parallel superposition
flow.
5.7.2 Theoretical Considerations 65 ,257
In the case of the superposition of orthogonal shearing in direction 3
on the main shearing in direction 1 (Fig. 4.1), the kinematic matrices

100

.-

-('Ij

p..

3'
10

CII

e,:,

-1

SHEAR RATE (s )
0 1: 0
2: 25.4
V 3, 3': 102
4, 4': 407

...

10

100
CJ

1000

(s -1)

SHEAR RATE (s -1)


1: 0
2: 25.4
V 3, 3': 102
... 4, 4': 407

.-

en

~3~__~~__~~~~

~CII

f,::"

...

4'

...

0.1 L-________

__________

________

100

CJ

(s

-1

1000

Fig. 5.19. Storage modulus G 23 and dynamic viscosity '123 versus frequency of
oscillations OJ at various steady shear rates Yo for solution 1 (Table 5.2). Lines are
theoretical predictions, symbols are data from Ref. 256.

146

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

100

10

SHEAR RATE(s -I)

1: 0
2: 25.4
V 3, 3': 102
"" 4, 4;: 407

10

100

(,.) (s
10

-1

1000

r-----------,------------.-----------,
1

....

'00

.C"d

p..

t'"l

4'

.N
~

4""

"" ""

SHEAR RATE (s
o 1: 0
2: 25.4
V 3, 3': 102
"" 4, 4': 407
0.1 L -__________
1

.l

-I

__________

10

(,.) (s
Fig. 5.20.

-1

100

__________

1000

The same as in Fig. 5.19 but for solution 2 (Table 5.2).

147

Shear Deformations in Elastic Polymeric Liquids

100
C\S

0..
6
M

C\l

SHEAR RATE (s

1:
2:
V 3:
4,
0 5:
6,
A- 7:
.Ii. 8,

10

"

10

100

-1

0
6.36
12.7
4': 25.4
50.9
6': 102
204
8': 407
1000

10
6'

.......

00

-C\S

0..

M
C\l

A- AA- A.Ii.

.Ii.

.Ii.

.Ii.

SHEAR RATE (s -1)

1: 0
2: 6.36
V 3: 12.7
" 4, 4': 25.4

5: 50.9
6, 6': 102
A- 7: 204
.Ii. 8, 8': 407
0

10

CJ

Fig. 5.21.

(s

-1

100

1000

The same as in Fig. 5.19 but for solution 3 (Table 5.3).

148

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

are of the form:


.1
e=2

Yo

[0~o

0
Ii

.1
W=2

[0~o
.

-Yo
0
Ii

~'l

c=

[e u

C12

C12

C22

C13

C23

en]
C23

C33

(5.49)

To simplify the notations, only one one relaxation mode is considered


here. Consider how eqns (1.9a) and (1.55) with kinematic matrices (5.49).
Because detc= 1, the matrix c- 1 is represented as:

-1

=[

C22C33-C~3

C13 C23- C12 C33

C12 C 23- C 13 C 22]

C13C23-C12C33

Cll C33-cI3

C13C12-C~lC23

C12 C23- C13 C22

C13 C12 -Cll C23

Cll C22- C12

(5.50)

Equations (5.49) and (5.50) yield:


11

=trC=Cll +C22 +C33

12 =trc- 1 =Cll C22 +C22 C33 +C33 C ll

-ctz -d3 -cI3

(5.51)

Then the total set of evolution equations for the components of tensor c
has the form:

2e

dC22

Tt +C12 +C22 +C23 +C22(I2 - I d/3-1 =0

2ed~~3 +cI3 +C~3 +d3 +C33(I2 -11)/3-1 =4C23eli

2e

Tt +(C22 +C33)C23 + C21 C31 +c23(I2 -

dC23

1 1)/3 = 2C22eB

2e

d~~l + (C33 + Cll)C31 + C23C12 + C31 (I2 -

1 1)/3 = 2e(C12 Ii

+ C23]i)
(5.52)

Equations (5.52) form the closed set which have the first integral:
detc= 1. Consider now the asymptotic solution for the superposition

Shear Deformations in Elastic Polymeric Liquids

149

of small amplitude-orthogonal oscillations with the (complex) shear


rate
8 = iwe exp(iwt)

(S.S3)

(Ie/ ~ 1)

on a steady shear flow with constant shear rate Yo. Then, C= Co + bc(t, e, Yo)
where Co is described by formulae (S.17) with Ilk = 1 and bc vanishes when
e--+O. In this case, 12 - II :::::;O(e) and the components C23 and C13 of the
order O(e), satisfy the following set of coupled equations:
20 d~~3 + (c? 1 + I)C13 + (C?2 - 2r)C23 = 2C?2 08

dC23 +C12
0
0 O'
20 dt
C13+ (0
C22+ 1)C23= 2C22
e

(r=yo)

(S.S4)

Searching for the steady oscillating solution in the form:


(C13, C23) =(gt3, g!3)e exp(iwt)

(S.SS)

we can obtain from eqns (S.S4) the expressions for dimensionless complex
dynamic moduli gt3 and g!3 of single Maxwell model as follows:

*_
2iwO[C?2(l +2iwO)+2rcg2]
g13 - (1 +C?l +2iwO)(1 +Cg2 +2iwO)+(2r -c?2M2

(S.S6)

*_
2iwO[cg2(1 + 2iwO) + 1]
g23 - (1 +C?l +2iwO)(1 +Cg2 +2iwO)+(2r -c?2M2

(S.57)

In the experiments,256 only the component of complex dynamic


modulus G!3 was measured, therefore attention is further focused on the
expression (S.S7). Using the formulae (S.17) for values c~, the real, g23,
and imaginary, g~3' parts of g!3 are represented as follows (see also Ref.
6S):
g23 =

Z+ 1 +2W 202
2W 20 2
Z Z2(Z + 1)2 + 4(z + 1)w 202 + 4W 4 04

- - -=---::-------=--=-----,-....,.

"
wO(z+ 1)
z+z2+2w 202
g23=
Z
Z2(Z + 1)2 + 4(z + 1)W 202 + 4W 4 04
(z= [(1 +J1 +4r2)/2r/2)

(S.S8)

These formulae are easy extended for the multimodal case with the
transformations: O--+Ok> r --+rk=OkY, Z--+Zk, g!3--+g:,23' Thus the

6.86% (by weight)


polyisobutylene in
cetane at 25C

8.54% (by weight)


polyisobutylene in
cetane at 25C

5.39% (by weight)


polyisobutylene in
cetane at 25C

Materials

No.

Table 5.2 Model parameters


tJ(Pas)
2.41 X 10- 1
6.22 x 10- 1
3.30 x 10- 1
1.23 x 10
1.40 x 10
4.51 x 10- 1
8.82 x 10
4.19 x 10
1.31 x 10

s
4.898 x 10- 2

3.152 x 10- 2

1.454 x 10- 2

2.54 x 10- 1
3.49 X 10- 2
4.54 X 10- 3

1.83 X 10- 1
2.13 x 10- 2
2.98 X 10- 3

2.65 X 10- 1
2.79 X 10- 2
2.71 X 10- 3

O(s)

256

256

256

Reference no.

151

Shear Deformations in Elastic Polymeric Liquids

dimensional dependences of G23 and 1123 = G23 /W on wand y have the


following form:
I

G23 (W, y)=2w


I '

n;' 11k ek

1 + Zk + 2w2e~

i" ~ z~(1 +zd 2 +4w2e~(1 +Zk +w2e~)

1123(W, Y)=11 SW +

11k(Zk + 1)
Zk

Zk + z~ + 2w2e~

z~(1 +Zk)2+4w2e~{l +Zk+w2e~)


(5.59)

Equations (5.59) demonstrate the anisotropy in the linear viscoelastic


properties of polymeric liquids induced by orthogonal superposition of
small-amplitude oscillations on steady shear flow. When y~O, eqns
(5.59) tend to the common formulae (5.21) of the linear theory of
viscoelasticity.
Equations (5.59) were compared in Ref. 257 with data obtained in Ref.
256. The comparison is shown in Figs 5.19-5.21. The parameters 11k, e k
and s for three-mode approximation were found in Ref. 258 by fitting the
standard dynamic data 256 for G/(w) and G"(w). These parameters are
represented in Table 5.2.
In Ref. 258, the calculations were made for the orthogonal superposition of small- and large-amplitude oscillations upon steady shear flow,
and for small-amplitude oscillations were compared with the data obtained in Ref. 256. It should be noted, however, that the calculations in
Ref. 258 were performed under the wrong assumption, C13 =0, which
contradicts the last (exact) equation in the set (5.52) and the asymptotic
(under e~O) first equation in the set (5.55) as well. It is because of this
mistake that the discrepancies between the calculations 258 and the
data 256 were so great.

5.8 SHEAR DEFORMATIONS UNDER GIVEN SHEAR


STRESS: START-UP FLOW AND ELASTIC RECOVERY
5.S.1 Experimental Results

Simple shearing under a given constant shear stress (when normal stresses
are usually varied) is commonly studied by applying a constant load to a
mobile surface of rheometers and measuring the motion of the mobile
surface. The experiments have been used in the parallel plate rheometer
(see, e.g. Ref. 259) and in the cone-plate rheometer (see, e.g. Ref. 156 and
Fig. 4.5(a)).

152

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Typical experimental plots 65 of shear strain l' versus dimensionless


time r: = t/B are represented by points in Fig. 5.22. Here l' = AI/H for
shearing in parallel plate and disk-disk visco meters, where H is the gap
width and Al the linear displacement; and 1'=t/J/rx for a cone-plate
device, where rx is the angle between the cone and plate and ' is the
twisting angle. The features of the time dependences of shear strain, y(r:),
and its time derivative, y(r:), at a given constant shear stress are revealed
in Figs 5.22 and 5.23. The shear strain 1'(r:) increases monotonously with
time while the shear rate first decreases and then, after passing a
minimum, increases tending to a constant value characteristic for steady
state flow. The minimum in the dependence y(r:) degenerates in the linear
region of deformations. The higher the shear stress, the faster is the flow.
In the steady state flow, the values of shear stress 0"12 and shear rate y
do not depend on which the constant magnitude, 0"12 or y, are initially
given.

12

8
o-{

/),-2

'l:=t/o

Fig. 5.22. Total shear strain y versus dimensionless time T = t/8 in start-up
shearing under given dimensionless shear stress ui 2 = u12/fJ,.65 Lines are theoretical predictions, symbols are the data for material 1 in Table 5.1. Numbers
correspond to Ui2 =0.86, 1.46, 2.07, fJ,= 2.12 x 10 3 Pa, 8=0.8 s.

Also, the elastic recovery shear strain can be measured in these


experiments (for a historical review, see Ref. 72). For this purpose, the
plate to be shifted (see Fig. 4.5(b is practically instantaneously unloaded
of stresses (0" 12 ~ 0), and then its reverse displacement is measured. As a
rule, the plate is allowed to move only in parallel to the other plate,

Shear Deformations in Elastic Polymeric Liquids

153

overcoming the resistance of descending normal stresses. At time


starting from the beginning of recovery, the value of elastic recovery
strain Ye(t) is defined as the ratio of the reverse travel of mobile plate to
the gap width H between the plates. The initial value Ye(O)=O and the
complete elastic recovery strain y", = Ye ( CIJ). The example of this dependence is shown in Fig. 5.24 according to Ref. 259.
8

o
Fig. 5.23. Theoretical dimensionless
time dependences of shear rate 1(1:)
under given constant shear stresses
with the values shown in Fig. 5.22.

Fig. 5.24. Time dependence of elastic


recovery strain Ye(t) ion the process of
retardation after steady shearing with
shear stress 0"12 = 104 Pa (fluid 2 in
Table 3.1).259

In start-up shearing with the constant strain rate 1, the complete elastic
recovery strain y",(t) depends on the time of shearing beginning with the
rest state. As shown in Fig. 5.25,259 this dependence, before reaching a
steady value, passes through the maximum. This plot is similar to the
dependence 0"12 (t) in start-up flow (see Fig. 5.4). The dependence of an
ultimate steady magnitude y", (00) on the shear rate, i.e. Y'" (1), increases
monotonously with 1 increasing as shown by the points in Figs 5.26 and
5.27,65.259.260 and may sometimes reach hundreds of per cent. 261 This
dependence is easier determined at steady flow (see, e.g. Refs 217, 227,
230,261-263). In Ref. 264 where the filled polymer was used, a maximum
on the dependence y", (1) was observed. When the values of 1 are not very
large, it is possible to estimate this dependence by the relation: Y'" (1) =
0"1(1)/(20"12(1)).1.72 When the values of 1 are very small, y",(1)=0"12/Ge
where Ge is the modulus of high elasticity. Some results were obtained in
Ref. 265 where the recovery strains were measured in the course of

154

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

three-dimensional unloading with both shear and normal stresses being


absent.
5.8.2 Theoretical Considerations 65 ,210
The shearing of elastic liquids under given (dimensionless) shear stress
O'r2(r) is under study in this section, Here the dimensionless shear rate
r(r) is unknown and should be determined, To simplify the theoretical
analysis, we consider first the set of constitutive equations (5.13) and
(5,14) with only one relaxation mode, which in the linear case degenerates
into the well-known three-constant rheological model (1,99), As follows
from relations (5,12), in this case, Ck=C and
vl=l-s,

f1 = 1](1- s)/()

(5,60)

Here 1], () and s(O~s< 1) are the rheological parameters of the model
which can be determined in the linear region of deformations, Introducing new variables and parameter m,
O'r2/S =Z(r),

C12(r)=m(z-r),

Cll(r)=y(r), m=s/(I-s) (5,61)

we can rewrite eqns (5,13) as follows:


2dy/dr+ y2 +m(r -z)[mz+(4+m)r] -1 =0

2dr/dt+(r _Z)/y[y2 +m2(r -z)+ 1] +2mr/y[(r _Z)2 + l/m2] =2dz/dt


(5,62)
The dimensional shear and normal stresses take the form:
0'T{r)=(I-s)[y2 -1-m 2(z- r)2]/y,
O'r2(r)= sz(r),
-0'!{r)=(I-s)[y-l-m 2(z-ry)2]/y

(5.63)

where the function z(r) is given. Two regimes of shearing are studied using
eqns (5.62) and (5.63), which are compared further with experiments.
(i) The start-up flow under given constant shear stress: z(r)=zoJ(r).
Here Zo = O'ho/s is a constant and J(r) is the Heaviside unit step
function. As follows from eqns (5.62), the initial conditions for the
functions r(r) and y(r) have the form:
y(O) = 1,

r(O)=z( +O)=zo

(5.64)

A qualitative analysis of the solution on the phase plane {r, y}


showed that when the value Zo is small, the function r(r) decreases
monotonously. When Zo is large enough, r(r) goes through the

Shear Deformations in Elastic Polymeric Liquids

155

minimum which is getting more intense and shifted to the region


of small r with Zo increasing.
(ii) Elastic strain recovery (retardation) after a shear pre-deformation.
In this case, at r=ro the mobile plate being suddenly released
from the shear stress is moving in the opposite direction to the
previous shearing (see Fig. 4.1). Only the case when the mobile
plate has a normal constraint and is forced moving in its own
plane is considered here. Under study are the time evolution of
normal stresses and shear rate l(r) due to eqns (5.62) and (5.63)
with z(r)::O when r>ro. The initial conditions for the set are:
l(ro +0) = c12(rO)/m,

(5.65)

Here cij(ro) are known values from the pre-shearing.


Investigation of the set (5.62) on the phase plane {r, y} shows
that the only steady point in the phase plane is the stable knot
{O, 1} which corresponds to the rest state; the solution being
monotonically decreasing over the entire time interval r > roo
In experiments, the elastic recovery strain Y(T) is commonly
measured, which is defined as the ratio of the plate pathway in the
direction opposite to pre-shearing, to the gap width between the
plates. This magnitude is written as follows:
(5.66)
The value Ye ( ex) ) = Y00 characterizes the total elastic recovery shear
strain.
Equations (5.62) and (5.63) with z(r)::O and the initial conditions (5.65) corresponding to the steady shearing (see eqns (5.27))
were solved numerically.
In Ref. 210, the retardation in the simple shear was considered
using the multimodal approach (N) 1) with s#O. The case s=O
was studied in Ref. 60. Though this approach insignificantly
changes the results, it leads to the analysis of jumps for the
components of elastic strain tensors Ck when the constant stresses
are suddenly applied to or are removed from the elastic liquid.
In order to determine the jump values, the formulae (5.30) are
used.
The theoretical predictions for both start-up flow with a given shear
stress (0"12 = constant) and retardation after steady shearing (0" 12:: 0), are

156

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

demonstrated by lines in Figs 5.22-5.28 in comparison with the


data. 65 ,259,26o The predictions of eqns (5.22) and (5.23) for y(r) and r(r)
are shown in Figs 5.22 and 5.23 for various constant values of 0"Tz, The
predictions for the time evolution of the first O"! and second O"~ normal
stress differences in the start-up flow with various constant values of
shear stress O"~ are depicted in Fig. 5.28 by solid and dashed lines,
respectively.

tUff J'

Fig. 5.25. Total elastic recovery strain Y00 versus total strain
Table 5.1 (Y=2s- 1 ).

yt

for fluid 2 in

-1

-2~__~~__L-~~_ _~~_ _~~~~~

-3

-2

-{

to/1

(5-1)

Fig. 5.26. Total elastic recovery strain Yoo versus shear rate y in preceding steady
shearing: 0, fluid 2 in Table 5.1; 1::., 57% PS solution in decaline at 26c. 259

Figure 5.26 and 5.27 demonstrate the theoretical plots (solid lines) for
total elastic recovery strain y", versus the shear rate y in the preceding
steady shearing. 65 ,21o Figure 5.27 also shows (by a dashed-dotted line)
the similar theoretical dependence C12(Y) for the fluid I (see Table 5.1). It
can be seen that the plots C12(Y) and y",(y) shown in Fig. 5.27 are very

157

Shear Deformations in Elastic Polymeric Liquids

f.S
o

o
/-

.,/'.

...

---

o "

Fig. 5.27. Theoretical (lines) and experimental (open circles; fluid 1 from
Table 5.1) dependences of total elastic
recovery Yoo on dimensionless shear rate
r in steady shearing. The dashed line is
the theoretical dependence of elastic
strain el2 on r.

O~----L...--

10-1

r
_ _---l
10

close to each other for moderate shear rates, i.e. the values Y00 measured
in the tests may often be taken for the values of elastic strains C12. When
values r are larger than those presented in Fig. 5.27, the dependence
Yoo(r) for N = 1 has a maximum, whereas the dependence C12(r) tends to
a constant value with r -+ ex).
5.9 ISOTHERMAL FLOW BETWEEN TWO ROTATING DISKS 65
5.9.1 Experimental Results
Consider the flow of an elastic liquid in the gap of width H between two
parallel and co-axial disks of radius R; the upper disk being rotated with
a constant angular speed n and the lower disk at rest (Fig. 4.7). In this
case, the shear rate y=nr/H is not uniform along the gap but increases
with the increase of radius r. The magnitudes of dimensionless mean
shear stress, Km and first normal stress difference, Kr averaged over the
radius are defined as follows:

(5.67)
where M is the torque, F is the axial thrust and f.L is a characteristic elastic
modulus whose value (along with the value of characteristic relaxation
time e) is presented in Table 5.1 for single relaxation mode (N = 1). The

158

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

K","

1
J

-2

to

~~~~~~::===!:~~~~~~
0 2 4
1:

Fig. 5.28. Dimensionless theoretical


time dependences of first atCr) (--)
and second a~(T) (---) normal stress
differences in start-up shear flow with
constant shear stress at2' Numbers 1,2
and 3 correspond to at 2 = 0.86, 1.46,
2.07, respectively.

1;
Fig. 5.29. Dimensionless torque Km
and thrust Kr versus dimensionless
shear rate r* in steady flow of Fluid 1
(Table 5.1) between rotating disks. 65
Lines, theoretical predictions; symbols
experimental data.

time dependences of Km(t) and Kf(t) in the start-up flow between


rotating disks is very similar to that of 0'12(t) and O'l(t) in uniform
shearing. The plots Km and Ke versus r* = QRe/ H are shown in Fig.
5.29 where various points correspond to different values of H =
2.5-4 mm. The plots Km(r) and Ke(r) in start-up flow between two
rotating disks for various constant values r*, and relaxation after
cessation of steady flow (r = 0), are presented in Fig. 5.30(a)-(c). In
transient flow, these dependences practically have no maxima which is
typical for 0'12 and O't in homogeneous simple shearing (see Fig. 5.4). This
is due to averaging the stresses over the radius. It is seen that normal
stresses K f reach the steady flow and relax somewhat more slowly than
shear stresses Km.
The shearing between rotating disks is much more convenient for
measuring time depending normal stresses than using the cone-plate
instruments because the effect of the lubricating layer near the cone
results in great errors (see Chapter 4). In the rotating disk device, the gap
between the disks can be changed without changing the nature of the flow

Shear Deformations in Elastic Polymeric Liquids

159

which appreciably decreases the effect of the lubricating layer on the


measurement errors for the transitional flows.
5.9.2 Theoretical Considerations
Using the correspondence qJ+-+1, z+-+2, r+-+3, between the cylindrical
coordinates {qJ,z,r} and the Cartesian coordinates {XhX2,X3} for the
simple shear, we obtain again eqns (5.13) with Yk=O for the evolution of
tensors Ck where the parameter r=nrO/H is now variable over the radius r.
Also eqns (5.14) for the stress components are valid here. Two quantities of
interest here are: the axial force F and the torque M defined as follows:
(5.68)
If necessary, the radial distribution of pressure p can be found from the
equation (see eqn (A.69)):

(5.69)
Inserting eqns (5.14) into eqns (5.68), after simple manipulations, yields:
Km=

sr4*:L VkAk+ r1 :LVk


3
k
*k

Jr.

ck,12(r,

nr 2 dr

(5.70)
where r * =nRO/H is the dimensionless shear rate at the edge of disk. In
the case of steady shearing between two rotating disks, the explicit
expressions are obtained after substituting eqns (5.33) for Ck,ij into eqns
(5.70) and following integrating. These are of the form:

sr * ~

1 ~

Vk

K m =4L.. V k Ak+ 24r 3 L..13(Xk- 6A k r *-1)


k

* k Ji.k

(5.71)

For the start-up flow and relaxation after cessation of steady flow, the
theoretical dependences Km(r) and Kr(r) were calculated numerically

160

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

using eqns (5.13) and (5.70). In these calculations, the initial conditions
(5.26) and (5.28) for start-up flow and relaxation, were respectively
used. The theoretical predictions for the fluid I (Table 5.l) are shown
in Figs 5.29 and 5.30 by solid lines for the first approximation and by
dashed lines for the second approximation. They are in fair agreement
with the experiment. The comparison between theoretical predictions and
4)

fl.t
000000

0,. ...... - - - - - - - - - - - - - - - - - - - - -

K... KI

Il.fJ.f

o /"

;y

</

4-'

"

Jf

/({

/;

~-

LI

_-----7J.---I::.--~-~-~-4
LI LI
t
0-1
1>-

it

C)
fl.t

---_____ 1fl.f

0-1
1>-2

Fig. 5.30. Dimensionless time dependences of torque Km(r) and thrust K r(,) in
start-up flow between rotating disks and subsequent relaxation (fluid 1, Table 5.1)
at various constant shear rate r*=(a) 0.312, (b) 0.496, (c) 0.785. 65 Lines,
theoretical predictions; symbols, experimental data.

161

Shear Deformations in Elastic Polymeric Liquids

data obtained here, supplements the results obtained for the simple
homogeneous shearing, since, in contrast to the latter, in shearing between rotating disks, both normal stress differences contribute to the
thrust.

5.10 ON CALCULATIONS OF SHEAR DEFORMATIONS USING


GENERALIZED POWER ELASTIC POTENTIAL
The use of the power elastic potential (eqn (5.7)) with n # 1 does not lead
to qualitative changes in the results obtained. Thus, with the given
kinematics of shearing and disregarding the stress distortions of potential
barriers (which, as shown in Chapter 6, is possible for simple shearing),
eqns (5.4) and (5.5) for the evolution of elastic strains coincide with eqns
(5.13) valid when 'l'k = 0 independently of the value of n. However, the
qualitative results of calculations presented in this chapter may change
drastically when using arbitrary values of n # 1.
In order to evaluate the effect of parameter n on the behaviour of
stresses, let us consider the case of steady simple shearing of a single
Maxwell model. From eqns (5.4) and (5.5) with}' =0, it follows: 38 ,39
cos 2<p =(c-c-1)/(c+ c- 1),

sin 2<p = 2/(c+ c- 1),

c2 _c- 2 =4r
(5.72)

which c is the principal value of elastic strain. Then the dimensionless


shear and normal stresses scaled by the value of elastic modulus jJ., are of
the following form:
Ot2

a~

=(l/n)(c n-

c-n)/(c + c- 1 ),

= (1/2n)(c n+ c- n-

2-

nan

aT = aT2(c - c- 1 )
(5.73)

It should be noted that unlike expressions (5.73) for the at2 and at, the

magnitude a~ is not invariant under transformation n--+ -no Thus we


should consider the cases of positive and negative values of parameter n
separately. The last formula in eqns (5.72) yields:
(5.74)
Then eqns (5.73) result in the asymptotic expressions for stresses for small
and large values of r:
(r ~ 1)

(5.75a)

162

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Zp-l/2
~
a*l2~~'
-a~~

1-2zP 2p

zP
2p

a!~l

(n>O),

a~ ~ 2a!(n < 0)

(p= Inl/2, z=4Irl)

(5.75b)

A qualitative comparison of data for simple shear and simple consideration of stability allows us to establish useful constraints on the values of
parameter n. First, experimental data show that the second normal stress
difference a~ is always negative and its modulus is considerably less than
a!. Due to eqns (5.75), this fact can be described if n>O, n<2. Second,
according to asymptotic expression (5.75b), the function a!2(r) increases
with increasing r if In I~ 1. Otherwise the function a!2(r) goes through a
maximum and its decreasing branch gives rise to the familiar onedimensional instability in shear flow. Thus, to describe the data qualitatively, the power elastic potential (1.51) is subsequently considered with
the constraints:
(5.76)
imposed on the possible values of parameter n. As shown in the next
chapter, for most studied polymers, n ~ 1.1-1.2.
5.11 QUALITATIVE FEATURES OF SIMPLE SHEARING AS
PREDICTED BY SOME OTHER MAXWELL-LIKE
CONSTITUTIVE EQUA TIONS 266
In this section, we briefly analyse some qualitative features of simple
shear flows predicted by Giesekus ll ,l2 and Larson's8o viscoelastic constitutive equations, as well as by an extended specification (1.56) of our
general viscoelastic model discussed in Chapter 1. In the following, only
the single Maxwell models are considered which, to unify the analysis, are
written in the canonical form represented in Section 3.4 (Chapter 3).
5.11.1 Giesekus model ll .l2
For the simple shearing, the general constitutive Maxwell equations (3.1)
and (3.6) with the specifications (3.45a), take the dimensionless form:
dCl2/dT + PCl2(Cll +C22 - 2) + C12 = rC22
dCll/dT + P(cIl + cI2 -2Cll + 1)+ Cll -1 = 2rC12

163

Shear Deformations in Elastic Polymeric Liquids

dC22/d't"+ f3(C~2 +ci2 -2C22 + 1)+c22-1 =0,


('t"= t/(J,
r=(Jy,
O~f3~ 1)

(5.77)

a~=a2/J-L=c22 -1

(5.78)

where J-L is the elastic modulus, (J is the relaxation time and 13 is a


numerical (fitting) parameter characterizing the flow-induced anisotropy
in viscosity. It is commonly known (see, e.g. Ref. 6) that when 13 = 1/2,
eqns (5.77) and (5.78) are reduced to the single mode constitutive
equations (5.13) and (5.14). This holds only for the simple shear flows.
According to eqns (5.77) and (5.78), the steady-state expressions for the
stresses at small and large values of r are of the following form:
(5.79a)

aT 2 ~ [(1 - f3)f f3r/ 2,

aT ~ [4(1- f3)/f3 3r/ 4r 1 / 2


-a~ ~ I-J2[i[(1-f3)ff3)]3/4r- 1 / 2
(r~ 1)

(5.79b)

The dependences aT2(r) for various values of 13 are represented in Fig.


5.31. When f3>!' the steady flow curves go through the maximum value,
max aT2 (13) = 1/(213) at r m(f3) = 1/(213 _1)2. In the inertialess approximation, the start-up flows with given r = constant and initial values
10,-~---,------,------.------~

1:\3 =0.1

;:~ :~;
4J =0.7
=0.9

5:13

__

N
....

,.,

/ ,

___ o

~---------

!_____

___________________

.... -

1_:,.::.................................................................

2
3

;/,<:: ..................................................
.. .. 5

0.1 "---_ _ _-'--______- L -_ _ _ _---'-_ _ _ _ _ _--'


0.1
10
100
1000

Fig. 5.31.

Dimensionless flow curves U!2(r) for the Giesekus viscoelastic model


with various values of parameter fl.

164

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Cij It~ 0 = Dij' converge to the steady-state values for any values of rand /3
including the case /3>t. This is the simple confirmation that
Theorem II (Section 3.3) is valid. The start-up flow also demonstrates the
appearance of overshoots on time dependences of shear and normal
stresses, which become more intense and shift to r-+O with increasing r.
However, it is easy to find that the decreasing branches of flow curves
with /3>t are one-dimensional unstable (i.e. unstable against ID disturbances) when the momentum balance equation is considered exactly,
including the inertia term. Some of these features of the Giesekus
equation are discussed in Ref. 267.
Consider now, using the inertialess approach, the start-up flow under
given constant shear stress O'TZ=C12 in the particular case /3=t for the
shear Giesekus equations (5.77) and (5.78). As mentioned previously,
these equations correspond to the single mode equations (5.13). As shown
in Section 5.8, when C12 < 1, the transitional flow is converged to the
steady shearing. Now we consider the case when C12 > 1. In this case, the
exact solution of eqns (5.77) with /3=t is of the form:
(0~/3~ 1),

c12(r)=~J(r),

Cll =(2 + aZ)jczz(r)

1- a tan (ar/2)
cZ2(r)= 1 + (l/a) tan(ar/2)'

a 2 =ci2-1

r(r)=c12D(r)+!~[(2+a2)/c~2(r)+ 1]

(5.80)

where J(r) is the unit step function, D(r) is the Dirac delta function and the
initial jumps of variables C12, Cll and r(r) have been taken into account.
Equations (5.80) demonstrate the 'blow-up' instability when the solution
goes to infinity at a certain finite time and immediately after that the
tensor c and the dissipation lose their positive definiteness.
Consider now the start-up shear flow under a given constant shear
stress C12 for the general case of Giesekus equations (5.77) and (5.78). Let
us rewrite the third equation for the component C22 in the form:
S=C22 -1 + 1/(2/3)

(5.81)

Equation (5.81) shows that if C1Z < 1/(2/3), the solution of the initial problem
tends to a certain steady value,
It has been shown 266 that in this case,
the transitional flow has features similar to those described in Section 5.8 of
this chapter. When C12 > 1/(2/3), the solution of eqn (5.80) is of the form:

'00'

, _ a 1- 2a/3 tan (a/3r)


2a/3 + tan (a/3r) ,

(5.82)

Shear Deformations in Elastic Polymeric Liquids

165

Though when /3#1 the complete analytical solution is not found, its
features are very similar to those described by eqns (5.S0). The blow-up
instability followed by lost of positive definiteness of tensor c and
dissipation also appears here.
5.11.2 Larson Model 80
For simple shearing, the general constitutive Maxwell equations (3.1) and
(3.6) with the specifications (3.43) and (3.44), take the dimensionless
canonical form:
dC12/dr + B(I 1)C12 = r,
dCll/dr + B(I l)(Cll -1)= 2rC12,
C22=C33 = 1,
B(Id= 1 +OC(Cll -1)/3
(JT2 =- (J12/ /1 = C12/B(I 1),
(JT =- (Jd /1 =(Cll -1),
(5.S3)
(r=Oy,
r=t/h,
O~oc~ 1)
where /1 is the elastic modulus, 0 is the relaxation time and oc is a
numerical (fitting) parameter characterizing the extensibility of strands.
Note that in this model, the second normal stress difference is equal to
zero.
In the steady shear flow, eqns (5.S3) are reduced to the following:
cll=1+2cI2'
C22=C33=1,
r=c12(l+2occI2/3)
OCT2 = C12/(l + 2occI2/3),
(JT =(1 + 2cI2)/(1 + 2occI2/3)

(5.S4)

When oc>O, the flow curves (JT2(r) go through the maximum with the
value max(JT2=J3/(Soc) at r=J6Ft (see Fig. 5.32); the first normal stress
difference monotonously increasing up to the ultimate value 3/oc when
r -+00. The inertialess analysis made due to eqns (5.S3) for start-up flow
with a given constant shear rate r on the phase diagram (e12, Cll) shows
that the only stationary point on the diagram is the stable focus. This
stationary point corresponds to the steady-state flow. It means the
appearance of overshoots for shear stress and the first normal stress
difference in the transitional flow and its converging to the steady-state
flow for any value of r. When using the complete calculations involving
the inertia term in momentum balance equations, the start-up flow is
one-dimensional unstable if the value of r relates to the decreasing
branches on the flow curves.
Also, the start-up flow under a given constant shear stress (JT2 = b was
analysed due to eqns (5.S3) in Ref. 266. It was shown that for positive
values of parameter oc, a transition to the steady flow exists if b < max (JT2 =
J3/(Soc). If J3/(Soc) < b < J3/(4oc), the blow-up instability occurs but is

166

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

1-0.=01
2-0.=0'3
3- (1=0'5
4- 0.=07
5- 0.= Og
6- 0.= 1'0

2
3

4
5
6

10

100

100C

r
Fig. 5.32. Dimensionless flow curves a!2(r) for the Larson viscoelastic model
with various values of parameter IX.

different to that occurring in the Giesekus equations. If b>J3/(41X), the


solution of eqns (5.83) does not exist.
5.11.3 Discussion of Blow-up Instabilities in Giesekus and
Larson Constitutive Equations
All the unpleasant instabilities found in single mode Giesekus (including the case P=! of single mode equations (5.13 and Larson constitutive equations have no physical sense and happened only because
of boundedness of shear stress in steady shear flows at r -+ 00. These
are not of a fundamental nature but are related to poor modelling of
the dissipative term at high values of De. These instabilities can be
delayed when using the multimodal approach but then one can have
some problems in two- and three-dimensional high De flows near the
corners and borderlines where the stress singularities may occur (see
Chapter 8).
In the Leonov model (P=!), these instabilities can easily be eliminated
by using the more complicated eqns (5.4) and (5.5) with y=O and value n
subordinated in eqns (5.76). Indeed, in this case, asymptotic expression

167

Shear Deformations in Elastic Polymeric Liquids

(5.75b) for O"T2, shows that shear stress is strictly growing when r~ 1.
Another possible way of providing the total shear stress with unlimited
growth is to add the viscous term to the viscoelastic shear stresses. If
the viscosity in the additional term is chosen properly, it is also
possible to eliminate the decreasing branches on the flow curves in
the Giesekus (/3>t) and Larson (IX>O) models and completely stabilize
the shear flow. If the viscosity in the additional term is small enough,
the decreasing branches of the flow curves in the Giesekus and Larson
models are not stabilized completely, and some windows of shear
instability will appear there. Note that the use of stabilizing viscous
terms completely changes the qualitative predictions of the viscoelastic constitutive equations in rapid flows and quick (jump)
deformations.
5.11.4 A Simple Viscoelastic Maxwell-like Constitutive Equation
This model uses the Mooney elastic potential (1.46) and specification
(1.56) for the irreversible strain rate with the power-like dependence of
relaxation time on the second invariant of elastic strain. The single mode
formulation of the constitutive equations is of the form:

t+ (/ 2/3)" [(l2/3)c - t5]/O = 0,

0"

= - pt5 + J.l[(1-IX)C-IXC- 1]
(5.85)

W=(J.l/2)[(1-1X)(l1 -3)+1X(l2 -3)]

Here, along with usual dimensional parameters (relaxation time 0 and


elastic modulus J.l), two numerical parameters, IX(O ~ IX ~ 1) and n, are
involved. This model is Hadamard stable and has limited values of tensor
C in any regular flow.
Equations (5.85) do not allow the planar deformations for the elastic
strain in simple shear, which is why the matrices c and c- 1 have the form:

~_,]

(5.86)

C3

Here the incompressibility condition


(5.87)
for the tensor c has been used. In this case,
(5.88)

168

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

and the kinetic equations for the evolution of tensor care:


dCll/dr + (/2/3t(CllI 2/3 -1)=2rC12
dC22/dr + (/2/3t(c22 12/3 -1)=0

(5.89)
The set (5.89) has the first integral (5.87) corresponding to the incompressibility condition. The system of dimensionless stresses scaled by the
modulus Jl is:
O'~ = -ctcr2

(5.90)

It is seen that the second normal stress difference appears only if ct > O. In
a steady-state flow, eqns (5.89) reduce to the set:

(5.91)

(z=I 2 /3)

10

2
3

0.=01

n = -2
2 - n= -1
3- n = 0

1-

4- n

= 1

5- n = 2

01

100

1000

Fig. 5.33. Dimensionless flow curves a!2(r) with various values of parameter n
(oc=O'l) for a new viscoelastic model.

169

Shear Deformations in Elastic Polymeric Liquids

Though there are no analytical solutions, the following asymptotes can


be easily found for the stress system at small and large values of r:
(r~

1)

(5.92a)

Ut2 ~(1- a)rl/(2n+ 5),

(r~

1)

(5.92b)

Asymptotic formulae show reasonable qualitative behaviour for small


and large values of r if n> -1. Dimensionless flow curves Ut2(r) are
shown for several values of n, a=O.l in Fig. 5.33. These are strictly
increasing, thus there is no instabilities of the types discussed in this
section. The calculations of start-up flows with a given shear rate
r = constant show overshoots in time dependences of shear stress and
first normal stress differences.

CHAPTER 6

Experimental and Theoretical Studies of


Uniaxial Uniform Extension of Polymeric
Liquids
6.1 INTRODUCTION
The present chapter deals with experimental and theoretical investigations of the simple elongation of polymer melts under isothermal conditions. This kind of deformation is the keystone in polymer processing
of fibres and films. To simplify the analysis, the deformation of cylindrical samples of high-viscous polymer melts is considered using inertialess approximations with kinematics given by eqns (4.34H4.37).
The schemes of common experimental devices are shown in Figs 4.10
and 4.11.
In view of the nonlinear properties of polymeric liquids considered in
Chapter 5, it can be expected that their rheological behaviour under
shearing and elogation conditions could be completely different. Besides,
in the case of extension, as distinct from shearing, there is no uncontrolled
interaction of polymer liquid with the wall from which it can detach in
high Deborah flows. This appreciably facilitates the studies of the
rheological behaviour of these liquids under very intense regimes of
deformation.
It is demonstrated below that the effects studied in this chapter are due
to finite elastic strains developed along with flow in simple extension.
Because of the simplicity of the deformation, there is also an opportunity
to subdivide reliably, the total strain into recoverable and irreversible

Uniaxial Uniform Extension of Polymeric Liquids

171

components. Thus, the investigation of the extension of elastic liquids


may yield substantial new data for understanding the nature of viscoelastic flow.
In order to test the methods of separating the strain into its constituent
parts, we consider first, simple extension in the region of linear deformations, which is reliably achieved at a constant linear rate at the end of
the extended specimen. These studies also give useful information about
basic rheological parameters of polymer melts.
In order to study the nonlinear behaviour of polymer melts, the
regimes of intense deformation under a given constant elongation strain
rate, stress or force are used, as has been discussed in Section 4.5. These
experimentally feasible regimes supplement each other well. They also
allow relatively easy theoretical treatment and occur in various polymer
processing operations.

6.2 STUDIES OF REGIMES OF LINEAR EXTENSIONAL


DEFORMATIONS 186 ,187
In order to illustrate the expediency of certain definitions, we consider
here the extension of polymer melts in the context of their linear
rheological behaviour. In this case, the viscoelastic phenomena can easily
be analysed by using the general linear relations (1.86) and (1.87). In the
particular case of the uniform extension of a viscoelastic cylinder, the
radial component of stress tensor a rr on the free surface of a cylinder is
equal to zero and the strain rate tensor is given by eqn (1.36) where
K = ezz Then eqn (1.86) for the longitudinal components of stress tensor
a ( = a zz) takes the form:
a=3

Ioo

(6.1)

m(t-r)K(r)dr

This equation is different only by factor of 3 from the respective equation


for simple shear, and therefore contains no new rheological information
on the linear viscoelastic behaviour.
Following the sense of linear viscoelastic relations (1.92), (4.40) and
(4.42), we assume that
a(t) 1
ee(t) = 3G o +3

ft_

00

m(t - r)d{r) dr

(6.2)

172

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where Be is the elastic (recovery) strain, m(t) is the creep function and an
overdot represents a time derivative. The first relation in eqn (6.2)
corresponds to the 'Trouton law'.268
When the recoverable strains are small,
(6.3)

Hereafter 1,10 and Ir are actual, initial and residual lengths of the stretched
sample. Equation (6.3) follows from eqn (4.40) when lell ~ 1. Note that
this condition can be satisfied even if the total strains are very great, i.e.
B = 1110 ~ 1. This means a predominant development of irreversible strains
in slow viscoelastic flows. When eqn (6.3) holds, the kinematic relation
(4.42) takes the form:
dBe

dt+ep=K

(6.4)

Now we consider the start-up elongation flow when the experimental


scheme shown in Fig. 4.10 is used and the free end of polymer cylinder
is extended with a constant linear velocity v = Vo. In this case, the strain
rate is represented as follows (see eqn (4.44)):

K(t) = J(t)Ko/(l + Kot),

(6.5)

where J(t) is the unit step function. This case of deformation allows us to keep
safely to the linear region of viscoelastic behaviour since K-+O when t-+oo.
Inserting eqn (6.5) into eqn (6.1) yields:

a(t) = 3Ko

f~ m(t -

r)dr/(l + Ko r)

(6.6)

The geometrical parameters of a stretched specimen are changed as follow:


1= 10 (1 + Kot/l),

R(t) = Ro(1 + Kot)-1 /2

(6.7)

where Ro is the initial radius of the specimen.


Using eqns (4.41) and (4.42) in addition to eqns (6.2), we can also
represent the relations for elastic Ie and residual Ir lengths of specimen in
the form:

(6.8)

Uniaxial Uniform Extension of Polymeric Liquids

173

In obtaining the first relation in eqns (6.8), we integrate by parts and


neglect the 'instantaneous' component of recoverable strain, eo = cr(t)!(3Go),
in the second eqn (6.2), because for polymer melts, the values of instant
modulus Go are very high (Go'" 109 Pa).
Consider now, the asymptotic behaviour of cr(t) and ee(t) when t-+oo.
In this case, the following formulae hold:

cr(t)= 31'/ Ko +3K2 M 1 + KoM2 +0([1 + K t] -3)


1+Kot
0 (1+K o t)2
0
(t-+oo)
ee(t) = m(t) {1

I'/K~

+Kot

(6.9)

+0([1 + Kotr2)}

where M 1 and M 2 are the moments of the relaxation function defined in


eqn (1.90), which are assumed to exist. The first relation in eqns (6.9) with
an allowance for the definition K(t) by eqn (6.5) demonstrates that in this
asymptotic case, the Trouton relation, cr(t) = 31'/K(t) holds. Equations (6.9)
also show that both functions cr(t) and ee(t) tend to zero when t-+oo; the
latter because m( (0) = G; 1, Ge being the high elasticity modulus. But due
to eqns (6.6) and (6.8), cr(O) = 0 and ee (0) = O. Thus each of the functions,
cr(t) and ee(t) have at least one maximum in interval [0, (0); the number of
maxima depending on the form of the relaxation function m(t). If each
function, cr(t) and ee(t) has only one maximum, crm = cr(t s) and maxee=ee(te),
then ts < te. This follows immediately from the expression:
dee I =!ftsdm(ts-r) dcr(r) dr>O
dt ts 3 0
dt
dr
where the integrand is positive. It is interesting to note that at the instant
te, where the maximum ee appears, the Trouton's viscous relation:
(6.10)
holds. This follows from eqn (6.4) and the first relation (6.2). Equation
(6.10) allows us to determine with a high accuracy the value of the
Newtonian viscosity 1'/ in the simple elongation experiment. Also, eqn
(6.3), the first relation (6.7) and the second relation (6.9) allow us to
establish a relation for the limit elastic recovery length le(t) when t-+ 00:
(6.11)
Equation (6.11) is also useful for the experimental determination of high
elastic modulus Ge or the characteristic relaxation time O.

174

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The problems of slow elongation flow under a given constant strain


rate" and constant given stress (T were also considered in Refs 186 and
187. In the steady elongational flow, when both these regimes of deformation coincide, the results are:
(6.12)
The formulae (6.12) allow us to determine the values of Newtonian
viscosity 1] and the characteristic relaxation time (J from experiments with
a slow steady elongation flow. These experiments were carried out in Refs
188 and 191.
Let us now describe the experiment where a constant linear velocity of
extension Vo of the free end of a specimen is given.186.187 The experiment
in the linear region of extension was carried out using PIB P-20 at 25C
with molecular weight M ~ 8.2 x 104 determined by its characteristic
viscosity in toluene solution. The Newtonian viscosity 1] and high elastic
modulus Ge were determined independently using a shear plastometer as
discussed in Section 4.2. It was found that 1] = 4 X 105 Pa s - 1 and
Ge =4 x 103 Pa. Two values of linear velocity vo, 1.93 x 10- 2 and 1.2 x
10- 2 cm s - 1, were used in the stretching experiments. The dependences of
recoverable lengths Ie are plotted versus time t in Fig. 6.1. These increase
monotonically and are saturated with increasing time. The value of high
elasticity modulus Ge = 3.9 x 103 Pa which was calculated using Fig. 6.1,

ie

(,.,.)

25
20

15
fO

5
0

90

ISO

270

a60

5'10

G30

t.C:s>

Fig. 6.1. Experimental plots of elastic elongation Ie versus time t in extension


tinder given constant velocity Vo: O. 1.93 x 10- 2 em s -1; D, 1.2 x 10- 2 cm s -1.

175

Uniaxial Uniform Extension of Polymeric Liquids

eqn 6.11 and the value of Newtonian viscosity (see below), is in good
agreement with the value of Ge found in the shear experiment.
Figure 6.2 demonstrates the time dependences of elongation stress O"(t)
and elastic strain 8e (t) = Ie/I, obtained in these experiments. Since each
experimental curve here has only one maximum, it is seen that the
predicted inequality ts < te is satisfied. Here ts and te are the instants
where stress 0" and elastic strain 8 e pass through respective maxima. Of
most interest was testing the constancy of viscosity '1 defined by formula:
'1 =!-[In(lr/lo)] -1

(6.13)

O"(r) d,

which follows from eqns (6.2) and (6.8). Within the limit of experimental
errors (5%), the value of '1 found from recalculating the data according to eqn (6.13), was independent of time and equal to 4.2 x 10 5 Pa S-1.
This value is also fairly close to that found in shear experiments. These
calculations speak in favour of defining the irreversible strain rate ep by
relation (6.4). The value of Newtonian viscosity '1 was also determined
using relation (6.10) and completely coincided with that found using
eqn (6.13). Thus the elongation regime Vo = constant was established

L--~--...I.-_--'-_---JL....-_--'-_...J

90

180

270

360

0.10

450 t(&)

Fig. 6.2. Experimental plots of stress IJ and elastic strain lell versus time in
extension under given constant velocity Vo: 0 and /::,., 1.93 x 10 - 2 em s - 1; 0 and
V, 1.2 x 10 - 2 em s - 1.

176

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

and approved as a simple and reliable method for determine the basic
rheological parameters rf and Ge in linear region of deformation.

6.3 EXPERIMENTAL STUDIES OF EXTENSIONAL


DEFORMATION IN A NONLINEAR REGION
Even the earlier experimental studies of extension flows of polymers 184 ,185,269 discovered that unlike Newtonian fluids, the ratio (JIK is
not a constant. Moreover, the relation (Jle p = 3rf, valid in the linear region
of deformation, was violated in the nonlinear region of deformation 184 ,185 where significant elastic strains accumulate in flow. Also, in
Refs 184, 185 and 269 it was reported that the ratio (JIK can be larger than
3rf, which, in contrast to simple shearing, demonstrated a certain stress
hardening under extension. It should be mentioned however, that in the
references cited, the experiments were carried out in non-steady and even
in non-uniform regimes of deformation. Therefore further experimental
works were aimed at careful studies of the increase in elongation viscosity
and the search for new possible effects in various regimes of extension. A
fairly comprehensive bibliography with discussions of the state of the art
in this field by the late 1970s is given in Ref. 270.
6.3.1 Evolution of Elongation Strain
(a) Extension under constant strain rate (K = constant)
In this kind of simple extension, the time development of stress (J(t) and
recovery strain &(t) were experimentally studied in many publications. 188 ,190,191,271-278 Some new effects discovered in this regime 190 ,273-278 are discussed in detail in the next chapter.
First let us dwell on the results of Ref. 271, where PIB (polyisobutylene)
P-20 at 20C (M ~ 105 , rf = 1.1 X 105 Pa s -1, Ge = 1.57 X 103 Pa) was used
as a test fluid. The parameters of the respective linear model (1.79H1.83)
were found in Section 5.5.1 as follows: the relaxation times 8 1 = 103 S,
8 2 = 10 s, the elastic moduli J.11 = 7.7 x 10 Pa and J.12 = 3.3 X 104 Pa. The
dependences of shear stress and the first normal stress differences on
shear rate yare shown by symbols in Fig. 5.8. Also shown by symbols are
all the experimental data below. In Figs 6.3 and 6.4, the dimensionless
ratio (J1(Krf) and recovery strain & versus dimensionless time 1: = tl8 1 are
plotted. The maximum value of & attained in these experiments was equal
to 15. When values of the elongation rate K are small enough, the ratio
(J1(Krf) does not depend on K at any time t, and for the steady elongational

177

Uniaxial Uniform Extension of Polymeric Liquids

G
K7

.9

.-1

v-2
0-3

b.-If

0-5
0- 6

J
1
10-2

10-(

Fig. 6.3. Time dependences of effective extensional viscosities in extension


under given constant dimensionless strain rate r = KO 1. Symbols 1 - 6 correspond
to the values r = 0.384, 1.2, 3.84, 12, 38.4 and 100, respectively. The solid and
dashed lines show the theoretical predictions with n = 1 and n = 1.15, respectively.
The experimental results are shown by symbols.
!It -

v - 2
A -

3
Ii

0-

0-

0-5

<0..
Fig. 6.4. Recovery strain & versus time t at various strain rates r. Theoretical
predictions (n= 1) are shown by lines, experimental data by symbols. Points 1-6
correspond to r = 0.384, 1.2, 3.84, 12, 38.4 and 100, respectively.

flow is equal to 3 (see the low curve in Fig. 6.3); the time taken to reach
the steady flow being of the order of 8 1 This is the region of linear
viscoelastic deformation where, in accordance with eqns (6.1) and (6.2),
the time dependence of the ratio a/(1]K) is three times higher than that in
the simple shear, a 12/(rrY).
With higher values of K, the time dependences of a/(1]K) coincide with the
linear dependence shown in Fig. 6.3 only in the beginning of stretching,
where the elastic strains are small enough. With increasing time, the sooner

178

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

the curves deviate from the linear one, the greater the value of K. At large
values of time t, steady-state flow can be reached, where time-independent
recovery strain ( and ratio a/(1]K) usually grow with K. In this case, the ratio
is commonly called effective elongation viscosity, 1]el. When values of K are
large enough, it is impossible sometimes to reach steady-state elongation
flow, mainly because of the limitations of the experimental procedure,
which are discussed in Chapter 4. For example, though no steady flows
were achieved at higher values of K in Ref. 190, they were found in Ref. 275
when it was feasible to use higher elongation ratios.
When a steady elongation flow was attainable, the elongation
viscosity 1]el was either constant 188 ,189,278 or grew with increasing
K190 ,271,272.274,277 starting from the value of 31] in the linear region; to
the author's knowledge, this growth was first recorded in Ref. 277. The
highest increase in 1]et. from 31] up to 201], was observed in Refs 190, 274
and 275, where also, the decrease on elongation viscosity was reported at
a very high strain rate. The latter effects are discussed in more detail in
Chapters 7 and 12. It is common for the ratio of elongation to effective
shear viscosities to exceed more than an order of magnitude (see, e.g. Ref.
274). The problems associated with the existence of steady-state elongation flow and elongation viscosity along with broad characterization of
stretched polymers are collected in Ref. 270.
With few exceptions, most papers report a monotonic increase of
recovery strain ( with time in an elongation flow with K= constant; the
function dIn (/dt being a smoothly decreasing function of time (see e.g.
Refs 190, 191, 271, 276, 277 and Fig. 6.4. In these cases, the irreversible elongation rate ep(t) increases monotonically reaching the value
ep ( (0) = K in a steady-state flow. The larger the value of K, the higher is
the contribution of the elastic component in total strain. A distinctive
feature of the curves presented in Figs 6.3 and 6.4 is the monotonic
decrease of time needed to reach steady flow with an increase in K. This is
not the case for the experiments described in Refs190.192,276 (see Chapter
7). Also, in contrast to industrial polymers, monodisperse polymers
demonstrate almost linear viscoelastic stretching behaviour up to a
certain critical stress and then they are suddenly ruptured279-281 in
elongation flow.
In Refs 191, 275 and 277 the time dependences of a/e p are represented
along with those a/K. In the linear region of deformation, the function
a/e p as mentioned above, is constant and equal to 31], while in the
nonlinear region, it changes from the value 31] to a steady elongation
viscosity 1]el if steady flow exists. Sometimes, when the value of K is high

Uniaxial Uniform Extension of Polymeric Liquids

179

enough, the ratio aje p goes through a maximum as time increases. 275
References 191 and 277 report that this maximum was observed very
close to the linear region.
(b) Extension under constant stress (a = constant)
In this case, the total and recovery & strains are measured over time.

Typical examples of dependence (t), with a monotonically decreasing


time dependence of derivative K(t) = din ejdt, are represented in Fig. 6.5.
Usually, when t-HXJ, the steady elongation flow is achieved where
K = constant. The elastic strain &(t) increases monotonically with time and
acquires a constant value on reaching steady flow. The irreversible strain,
ej&, also increases with time, while the time derivative of its logarithm, ep '
tends to the constant value coinciding with K in steady flow. When
a=constant, the ratios ajK and ajep are monotonically increasing functions of time; the second dependence passing above the first one and
reaching the value 311 at t = O.
Comparison between steady extensions under conditions K = constant
and a = constant shows that they coincide with equal K and a.
(c) Extenion under constant force (F = constant)

This regime of simple elongation corresponds to stretching with intensively increasing stress, when the steady flow is not obviously attained
and deformation, sooner or later, leaves the linear region. The duration of
linear deformations is evaluated in Ref. 186.
In this regime, the time dependences of total strain e and its elastic
part & were measured for PIBs of various molecular weights and
LDPE. 184,185,193,259,282,283

o L.--fJ...O-...,,2o":--;r-=o:----:'#":O--:5~O,---'

t(s)
Fig. 6.5. Experimental plots of total strain time t in extension of polyethylene
(fluid 2 in Table 6.1, T = 150C) under given constant stress 0'0' Symbols 1 and 2
correspond to 0'0 = 8 X 104 Pa and 3.1 x 104 Pa. 273

180

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Figures 6.6 and 6.7 show the plots = 1/10 and eX versus dimensionless time
r = t/fJ l 193 obtained for PIB P-20 under constant force F for various initial
stresses (Jo=F/S o where So is the initial cross-section of the stretched
specimen (remembering that for the PIB, fJ l = 10 3 s and III =7.7 x 10 2 Pa).

e
1\

0-1

0-2

A-$

0-4
f01

Fig. 6.6. Dependences of total strain on dimensionless time , in extension


under various constant stretching forces (initial stress 0'0)' The lines show the
theoretical predictions; symbols 1-4 represent the experimental results with
dimensionless initial stress O'~ = 1.3, 4.29, 8.2 and 15.2.
1\

fX,

101

0-1
0-2
Il. - 3

If

00

Fig. 6.7. Recovery strain Ii versus dimensionless time , in extension under


various constant stretching forces (initial stress 0'0)' The lines show the theoretical
predictions; symbols 1-4 represent the experimental results with dimensionless
initial stress O'~ = 1.3, 4.29, 8.2 and 15.2.

181

Uniaxial Uniform Extension of Polymeric Liquids

In the time interval under study, the functions 8'(1") and &(1") increase
monotonically and are independent of (Jo at low values of 1". With time,
the larger the value of (Jo, the faster 8'(1") and &(1") increase.
For practical purposes, the dependence &(8') shown in Fig. 6.8 may
serve as a measure of macromolecular orientation for non-uniform
drawing, as in the case of fibre formation (see Chapter 10). In the regime
of extension with F = constant, this function is practically independent of
the value of initial stress (Jo. The case of extension of a pure elastic
material corresponds to a straight line &=8' drawn at 45, which is also
shown in Fig. 6.8.

/
:
Y-

0"

~~A

6. -

..

--3

..

x-I;

Fig. 6.8. Experimental plot of recovery strain eX versus total strain in extension
under various constant stretching forces (initial stress 0"0)' Symbols 1-4 correspond to the experimental results with dimensionless initial stress O"~ = 1.3, 4.29, 8.2
and 15.2.

Unlike shearing, the superposition of oscillations on simple extension is


very difficult to carry out experimentally. An example of such a superposition is discussed in Ref. 284.
6.3.2 Dependences of Stress and Strain Rates on Elastic
(Recovery) Strain
Now we shall dwell on regularities common to various regimes of
extension. The plots of stress (J versus recovery strain & are shown
in Figs 6.9 and 6.10.193.271 Open symbols in Fig. 6.9 denote the

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

182

&

(Pa.)

10'

(Pa.)

10 5

er:.'t

105

1>1>.16

"1
,.
",.

" ~""

o ._

'C"

...1'.

~.

- 1
"1-2

0-3

A -

4-

0- S
0- G

n.

1
2

-3
o - 46

10 3

Fig. 6.9. The plots of stress a versus


recovery strain & at various given
constant strain rates r. Theoretical
predictions are shown by lines, experimental predictions points. The
solid symbols denote the envelope of
Symbols 1-6 cordependence a(&,
respond to the values r = 0.384, 1.2,
3.84, 12, 38.4 and 100, respectively.

IJ [J -

Fig. 6.10. The plots of stress a versus


recovery strain & at various given
constant stretching forces. Theoretical
predictions are shown by lines, experimental ones by points. The solid
symbols denote the envelope of deThe figure designapendence a(&,
tions are presented in the caption
for Fig. 6.6.

n.

dependences 0"(&) in extension under various constant elongation rates.


These were obtained from the experimental data displayed in Figs 6.3 and
6.4. Values to the extreme right relating to maximum Ii correspond to
steady flow when it was attained. Figure 6.9 shows that the stress 0"
depends not only on Ii, as is often assumed, but also on the elongation
rate K. Figure 6.9 also demonstrates that the experimentally found

Uniaxial Uniform Extension of Polymeric Liquids

183

function, (J(K, &) is nonlinear in both variables, since the following


inequalities are satisfied:

The open symbols on Fig. 6.10 denote the dependences (J(&) obtained in
extension under given various constant stretching forces with initial
stresses (J o. These curves are plotted using the data presented in Figs 6.6
and 6.7. Figure 6.10 shows that the experimental curves (J(&) become
closer when & increases.
A comparison of dependences (J(&) obtained in regimes under given
constant elongation rates (Fig. 6.9) and constant force (Fig. 6.10) shows
that the strain rates are approximately identical at the points of their
intersection. To establish this fact the dependences n&) = 8 1 K(&) obtained
in extension with F=constant, were used (see Fig. 6.11). These curves are
plotted for various values of (Jo using the data shown in Figs 6.6 and 6.7.
The open symbols in Figs 6.12 and 6.13 denote the plots of the
dimensionless irreversible strain rate 8 1 ep (&)=r-dln&/dr versus & at
various values of given dimensionless elongation rate r = 8 1 K and force
(JO/J11, respectively. The functions &(r) shown in Fig. 6.4 were used for
constructing the curves ep (&) in the regime r=constant. The last point on
each curve at the highest value of & corresponds to a steady elongation
flow, if it was attained. In order to plot the curve ep (&) in the regime
F=constant, the experimental dependences (r) and &(r) shown in Figs
6.6 and 6.7 were used. As seen from Figs 6.12 and 6.13, the curves ep (&)
become closer when & increases.
The solid symbols in Figs 6.9, 6.10, 6.12 and 6.13 show the envelope
of the dependences (J(&) and ep (&), respectively. These envelopes are
obtained in Refs 193 and 271 from the stress relaxation experiments. In
the course of stress relaxation, the various dependences (J(&) and ep (&)
form these envelopes when the recovery strain &decreases. Steady values
of (J(&) and ep (&) in tests with r = constant lie practically on these
envelopes.
The existence of the envelope for the plots (J(&) allows us to find the
maximum elastic strain & attainable at a given stretching stress. This is
important for processing fibres and films since the strength of a polymer,
glassy after extension, increases with increasing &.
As an example, consider the simple extension under given constant
stress (J. In this case, the dependence (J(&) is the straight line parallel to
axis &. Its intersection with the envelope gives the maximum value of &
attainable at an assigned stress (J.

184

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

o
L

0- 1
0-2

to - 3

o
o

0-4

Fig. 6.11. Dimensionless


strain rate r versus recovery
strain Ii at various given
stretching
forces
(initial
stresses). Solid lines show theoretical
predictions;
the
points represent experimental
data. Symbols 1-4 represent
11~ = 1.3, 4.29, 8.2 and 15.2.

10

A comparison between the dependences (J(&) and e p (&) for various


regimes of extension, and stress relaxation after these, suggests 285 ,286 that if
for two processes, the curves 0"(&) intersect at a point &0, the curves ep (&) also
intersects at this point. This means that a functional dependence exists:
(J = (J(&, ep )

As shown in Section 6.2,

(J~31Jep

when

&~l.

(6.14)

185

Uniaxial Uniform Extension of Polymeric Liquids

....f
I"
~o~
o

10

17~

'"
10

111

",tI-

..

..

0- 1

2
3
0-4

o A -

10

v - 2
o - J

If
() - 5

tl -

"

o - {}

10 1 a:;
"

Fig. 6.12. Irreversible strain rate e p


versus recovery strain f2 at various constant strain rates r. Lines show theoretical predictions; points represent experimental data. For figure designation
see Fig. 6.3.

(0- 1
10

Fig. 6.13. Irreversible strain rate ep


versus recovery strain f2 at various constant stretching forces. Lines show theoretical predictions; points represent
experimental data. For figure designation see Fig. 6.6.

It should be also noted that at the points of intersection of dependences


0"(&) or ep (&) for two processes of elongation, the total strain e is
approximately identical and in active regimes of extension (K > 0) the
elongation rates K are identical too.

6.3.3 Stress and Strain Relaxation


The existence of functional dependence (6.14) suggests that after different
uniform extensions (and even after a preliminary partial stress relaxation),

186

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

the relaxation processes proceed in time identically (and independently of


deformation prehistory) if at the beginning of relaxation, both the irreversible elongation rate e p and the accumulated elastic strain & have the same
initial values for two processes. In accordance with this, instead of
parameters (&, ep ), one can take, for example, another couple (&,0) and,
for active extensions, also (K, a).
To test these findings the relaxation experiments 287 were carried out
using PIB at 20C, with properties somewhat different from the
previous one: '1:::::;1.3xl0 6 Pa s-l, Ge :::::;4.4xl0 3 Pa. The initial values
of magnitudes a, &, K and ep for the relaxation processes are given in
Table 6.1.
Table 6.1
The values of stresses and kinematic parameters at the initial moment of
relaxation processes (see Figs 6.14 and 6.15)
NN of intersection
points

The method
of their
attainment a

O'xlO- 5

I
II
II

1
2
3
4
5
6
7
8

1.35
1.35
1.50
1.50
0.45
0.45
0.27
0.27

III.

IV

IX

(Pa)

---at

KX 10 2
(s -1)

ep x 10 2
(s -1)

1.82
1.75
1.61
-1.63
0.30
-0.63
0.47
0.43

3.25
3.07
3.25
0.00
0.93
0.00
0.93
0.89

1.43
1.32
1.64
1.63
0.63
0.63
0.46
0.46

d In IX

x 10 2
(s -1)

4.0
4.0
4.8
4.8
3.0
3.0
1.9
1.9

'Points 1 and 3 are achieved in extension with K=3.76 x 10- 3 s-'; points 2 and 8, in extension with
(0'0 =6.3 x 10 3 Pal; points 4 and 6, during extension with K= 10- 1 and 3.25 x 10~2 S-1.

F=constant

The results of the experiments on stress relaxation are demonstrated in


Fig. 6.14, and that for the strain relaxation (retardation) in Fig. 6.15
where is the length of the specimen in course of retardation. In these
figures, the symbols 1 and 2 correspond to (a, &) with number I; symbols
3 and 4 to II; symbols 5 and 6 to III; and symbols 7 and 8 to IV.
Each symbol with its respective Roman numeral was obtained in experiments with two different prehistories of extension as explained in
Table 6.1.
As seen from the figures, the above statements are confirmed experimentally. Moreover, it was found that if two relaxation processes coincide
in time, then under the same initial conditions (&, ep ) the two retardation

r.

187

Uniaxial Uniform Extension of Polymeric Liquids

D-1~I
-2

'" -3 J/
" -II

O-s}
e-Ii 8/

(OIJ

(00

Fig. 6.14. Relaxation after extension up to a given stress


&: I-III correspond to Table 6.1.

(J

and recovery strain

1\

l,(tJ

D -

f}I

--2

,.-/13}11

('0

A-

)( -7}1B

.. -8

0.8

,..

0.2
0

-"" .,
D

Ii

600

iii

1200

III

(BOO

ttl)

Fig. 6.15. Time variation of relative length of a specimen in retardation experiment after extension up to a given stress (J and recovery strain &. I-III correspond
to Table 6.1.

188

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

processes proceed identically, automatically. It is also interesting to note


that at a fixed initial stress cr, the stress relaxation proceeds more slowly
for higher initial values of elastic strain, &.

6.4 CONSTITUTIVE EQUATIONS FOR UNIFORM EXTENSION


As shown in Section 4.5, in a cylindrical coordinate system, the kinematic
matrices are:
e=K diag{l, -1/2, -1/2},

(1)=0
(6.15)

where K is the extensional strain rate, which generally depends on time,


and the conditions of incompressibility (1.74) have been taken into
account. Substituting the matrices (6.15) into eqns (1.72H1.76) yields:
1 dAk (Ak+1)(A~-1)
Ak (it+
60~A~
-K,
(6.16)

Equations (6.16) are the set of constitutive equations for the simple
extension, which represent the relations between the external parameters,
elongation stress cr and strain rate K, and internal parameter A. In this set,
the specifications (1.58) or (1.59) of relaxation time O~ can be taken into
account. Also, in deriving the formula for the total stress cr, it was
assumed that the radial component of stress tensor is equal to zero on the
surface of the stretched cylinder, which means that the surface tension is
assumed to be negligible.
As a simple approximation of the set (6.16), consider now, the model
consisting of two parallel Maxwellian elements with neo-Hookean elastic
stresses, connected in parallel. 38 In this case, assuming that the relaxation
times Ok are functions only of temperature T, we have nk = 2 and s = 0,
and eqns (6.16) take the dimensionless form:
(6/x)dx/d't' +(x 2 _x- 1 )(1 +x- 1 )= 6r
(6/y)dy/d't'+(y2 - y-l )(1 + y-l )//3= 6r

(6.17)

Uniaxial Uniform Extension of Polymeric Liquids

189

Here the dimensionless variables and parameters are introduced as


follows:
Y=AZ'

v=J1,z/J1,l,

(6.18)
(P~

1)

and Ok and J1,k are the relaxation times and elastic moduli, respectively. In
the following, we consider various cases of simple extension (K~O) in
inertialess approximation which seems to be natural for the polymeric
systems under study.
6.5 THEORETICAL TREATMENT 38 - 40
6.5.1 Extension Under a Given Constant Strain Rate (K = Constant)
In this case, eqns (6.16) represent independent differential equations for
the parameters Ak and one algebraic relation for the stress (J. If the
start-up flow from the rest state is under study, the initial conditions are:
(6.19)
We consider in this section only the case when O~ = Ok> meaning that the
relaxation times are independent of elastic strains; nk having arbitrary
admissible values.
Being more interested in steady extensional flow, we can introduce the
constant dimensionless strain rate r k = Ok K for the kth relaxation mode.
If minkrk~ 1, the asymptotic representation Ak~ 1 +ek (O<ek~ 1), for
the elastic strains Ak is valid, which when substituted into eqns (6.16) gives
the linear (Trouton's) relation for viscous elongational flow:
(J

(6.20)

= 3IJK

If mink r k~ 1, the Ak ~ 1 and the asymptotic formulae for stress


elastic strains Ak are:

(J

and

(6.21)
k

Consider now, the dependence of elongation viscosity IJe = (JIK on strain


rate K in a steady extensional flow. If nk > 1, IJe increases unlimitedly with
increasing K and decreases at high values of K otherwise. If it is assumed
that all nk have the same value n independently from the relaxation mode,
then the dependence IJe(K) is consistent with the ineqns (5.66) established
for the shear flow (1 ~ n ~ 2) is described as follows:

190

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

If 1:::; n:::; 2, the elongation viscosity increases monotonically with


inc rasing K. If n = 1, it has the limit, 611, when K-HJ). If 1 < n:::; 2, it grows
unlimitedly with growing K. More discussions about the modelling of
elongation viscosity in comparison with the experimental data are given
in the next chapter.

6.5.2 Extension Under a Given Constant Force (F=constant)


Following eqns (4.38) and (4.46) discussed in Section 4.5, the stress in this
regime is given by

(J = (JoB

(6.22)

B= l(t)llo

Here l(t) and 10 are actual and initial (at dimensionless time r = 0) lengths
of a cylindrical specimen, So is its initial cross-sectional area, B(t) is the
total strain of the specimen and (Jo is the engineering stress. The
dimensionless extensional strain rate is represented by
(6.23)

r=d In B/dr =B- 1 dB/dr

Substituting eqns (6.22) and (6.23) into rheological equations (6.16) results
in a closed set of equations relative to variables Ak and B.
When, in eqns (6.16), s#O, the initial conditions for the set corresponding to the start-up flow from the rest state, are:
(6.24)

When s = 0, the elastic jumps appear in every relaxation mode, as well as


for the total strain. Thus the initial conditions at r = + change for the
following:
(6.25)
where Bo is an unknown constant searched for. In order to find the value
of this constant at the instant r = + 0, we should, using conditions (6.25),
write two expressions for elongational stress (J, following from eqn (6.16)
and eqn (6.22), and equalize them. In so doing we obtain the following
equation for the constant Bo:
(6.26)
In the particular case of simplified eqns (6.17), eqn (6.26) is reduced to
(1+V)B6+soB~=1+v

(so = (JolJ-ld

(6.27)

In order to understand the qualitative peculiarities of simple elongation under a given constant force, let us consider eqns (6.17) within a

191

Uniaxial Uniform Extension of Polymeric Liquids

so-called quick adaptation approximation 38 of the y-mode when the time


derivative in this second relaxation mode is neglected (13 ~ 1) and the
elastic deformation there is assumed as small (I y - 11) ~ 1). In this case, the
equation for the second mode is reduced to
(6.28)
and the dimensionless relation for stress in eqn (6.17) can be rewritten as
follows:
(6.29)
where
0" * = 8/(1 + vf3),

(6.30)

s=vf3/(l +vf3)

Along with the equation


r=O";1 dO"*/dr

(6.31)

which follows immediately from eqns (6.22) and (6.33), the first equation
in (6.17) and eqn (6.29) form a closed set relative to variables x and 0"*.
This set can be represented in the standard form as follows:
X-I

dx/dr = -(x + 1)(x 3 -1)/(6x2)- [(1- s)(x 3 -l)/x - 0" *]/(3s)

(6.32)

The initial conditions are:


(6.33)
The study of the set (6.32) on the phase plane {O" *, x} shows that the
phase solution O"*(x) decreases monotonously. This gives rise to the
asymptotic solution of the problem. Two asymptotes solutions are
obtained:
0"*:::::: O"~x,

::::::x:::::: 1 +0"~r/3s+ ...

0"*::::::x 2 ::::::3/(a-r)

(x-l
(x~

1)

1)

(6.34a)
(6.34b)

where a is a constant. Asymptotic expression (6.34b) represents a typical


blow-up solution, which reaches infinity in a finite time. It corresponds to
the evident physical sense as eventual rupture of a liquid specimen under
constant stretching force, if inertia forces are negligible. Nevertheless, in the
region of x commonly compared with the experiments, the inertia terms are
negligible and the intermediate asymptotic solution (6.34) is still valid.

192

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Matching both the asymptotic solutions at an intersection point om,


yields:
(6.34c)
Formulae (6.34) give an approximate qualitative description of simple
extension under given constant elongation force.
An example of time dependence of In e for the model considered is
represented in Fig. 6.16. When in the power elastic potential, n> 1, the
inflection point appears on the dependence In e(t) which disappears when
n-+1. Note that while in the regime K=constant, the variations in the
value of n affect only the values of stress force; in the regime of
F = constant, these variations change the course of the extension process
in time.
J,-------,------------n

tng

50

Fig. 6.16. Theoretical predictions for


time dependences of total strain (t) in
extension under given constant force
(0"0 = 104 Pa). The values of parameters:
(Jl =82 s,
/1=2.6 x 10 6 Pa, s=0.25.
Curves 1 and 2 correspond to n = 1 and
t(1)

n = 1.8.

6.5.3 Extension Under a Given Constant Stress (0' = 0' 0 = constant)


In this regime of extension, the variables K and Ak are unknown in the
set (6.16). The initial conditions (6.24) or (6.25) are also valid in this
regime, depending on value of parameter s. When s=O, eqn (6.26)
should be solved to determine the initial step-wise variation eo of strain.
When ,-+00, the set (6.16) tends to the steady equations coinciding with
those considered under given constant extensional rate K. Asymptotic

Uniaxial Uniform Extension of Polymeric Liquids

193

dependences for the steady elongational flow are represented by eqns


(6.20) and (6.21).
6.5.4 Elastic Recovery after Uniaxial Extension
Here we attempt to find the relations between the elastic strains Ak in
relaxation modes, and experimental values of recovery strain & and
irreversible strain rate ep defined by eqns (4.40) and (4.42), respectively.
Let us consider the retardation process (r>O) in a viscoelastic cylinder
that was previously (r < 0) extended with a given strain rate r and was
instantaneously unloaded at the instant r = O. At r = - 0, Ak = A~ and
(J = (J 0; these values being determined by previous deformation and
assumed to be known. Let 10 , I, Ir(r) and ~ (~=/r(oo)) denote the following
lengths of the sample: initial, actual at the instant of unloading, actual
length during recovery process and final length after complete recovery,
respectively. Then the obvious inequalities

(6.35)
hold true which reflect the active process of elongation and the liquid
viscoelastic nature of the medium. Of course, ineqns (6.35) also follow
from the set (6.17).
Introducing the notations
ex(r) = I/Ir(r),

&=ex(oo)

(r=t/Od

(6.36)

we can define the extensional strain rate r(r) during the retardation
process, as follows:

(6.37)
Thus, by taking into account eqn (6.37) and the fact that during
retardation, the total axial stress (J is equal to zero, we can rewrite eqns
(6.16) for retardation in the form:

Ok dAk (Ak+1)(A~-1)
Ok dex
-+
-=---

Ak dr

60~A~

ex

dr

(6.38)

The set of equation (6.38) should be complemented by some initial


conditions. When s0;60, the initial conditions are continuous:
(s0;60)

(6.39)

194

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

When s = 0, all the Ak have the same jump a - t, where a is an initial jump
in the retardation; the value of the jump determining from the condition
that at r = + 0, the stress (J is equal to zero. Thus the initial conditions are
of the form:
a( +O)=a,

(s=O)

(6.40)

with the second equation serving to determine the value a.


As an example of calculations of retardation process, we consider the
set of equations (6.17) where r= _a- 1 da/dr, 8=0 and x( -0)=
Xo, y( -0)= YO.38 The value of initial jump a is found from the equation:
(6.41)

given by:
a = [(x6 + f3Y6)/(xo 1 + f3yo 1)] -

1/3

Since f3~1, then xo>Yo, and therefore at r>O, we have x>1 and y<1.
This means that after instantaneous unloading, the x-strain remains
stretching whereas the y-strain changes from stretching to compressing.
A further solution of the problem requires numerical calculations.
However, it is possible to find an approximate analytic solution and
establish a useful relation between Xo, Yo and total recovery strain &
which can be checked experimentally. Consider the third equation in the
set (6.17), which in the retardation (8=0) is represented as follows:
(xo>x~1,

y<1)

(6.42)

This equation gives an one-to-one monotonously increasing dependence


y(x) which can be approximately represented in the region x ~ 1 with an
accuracy of not less than 3 % as follows:
3f3+X 2 _X- 1
y(x)=f3 3f32+(X 2 _X- 1 )(2f3+X 2 _X- 1 )

(6.43)

Consider now the phase diagram {a, x}. On the phase plane, we can
represent the set of equations in the form of one equation:
1 da
a dx

1 + y(x) + vf3(l + x)y'(x)


x[1 + y(x)] + vf3(1 + x)y(x) ,

where solution a(x) is found by simple integration.

(6.44)

Uniaxial Uniform Extension of Polymeric Liquids

195

When x = 1, which corresponds to the limit t-+ 00, the respective value of
IX is equal to 1, i.e. 1X(1) = 1. This value is calculated as follows:
A

lX=a exp

[fxoa1

l+y(x)+vfJ(l+x)y'(x) ]

x x[1 + y(x)] +vfJ(1 +x)y(x)

(6.45)

To calculate this integral we should substitute the function y(x) found


from eqn (6.42) (or approximate formula (6.43)); the jump value a being
calculated from eqn (6.41).
In experiments, the unloading procedure under retardation conditions
is often used repeatedly at different times of active stretching. In this case,
the dependences xo(t), yo(t), a(t) and 1(t) can be also obtained, and
using the latter, the irreversible strain rate ep(t) can be found from eqn
(4.42). Using a simplified, quick adaptation approximation (eqn (6.29)),
along with the first eqn (6.17) and the same procedure presented above
for two relaxation modes, the relation 1(x) can be easily derived and
represented as follows:
1(x)=zP(x),

z(x)=s/2+(I-s/2)x,

p=(1-s)/(I-s/2) (6.46)

Substituting eqn (6.46) into eqn (4.42) and using further the first equation
(6.17) yields:
81 ep= [sr + (l-s)(x 2 - x- 1 }/3](1 + x}/2z(x)

(6.47)

where z(x) was defined in eqn (6.46). Equations (6.29) and (6.47), defining
the elongation stress (T(x, r) and irreversible strain rate ep(x, r), result in
the fact that at various fixed values of r, these functions have lower limits
(T(x,O) and ep(x, 0), respectively.
6.6 COMPARISON OF THEORETICAL CALCULATIONS
WITH DATA
The above theoretical predictions have been compared with experimental
results on a simple extension of PIB p_20. 38 ,285 For this reason, the
general constitutive eqns (6.16) were solved numerically under assumption 8~ = constant for various regimes of start-up flow: (i) r = constant
with initial conditions (6.19); (ii) F=constant allowing for eqns (6.22) and
(6.23), with initial conditions (6.25) and parameter eo determined from
conditions (6.26) and (6.27); (iii) for the quick adaptation approximation
(6.29), where the quantities & and ep were also calculated using eqns (6.45)
and (6.32).

196

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The four independent rheological parameters of the model were chosen


for the PIB on the basis of the shear experiments discussed in Chaper 5.
The values of the parameters are:
111=7.7 x 10 5 Pas-1,
/11 =

7.7 x 10 Pa,

112=3.3x 105 Pas- 1


/12 =

3.3

104 Pa)

In the case of quick adaptation approximation (6.29), only the three


parameters, 8 1 =10 3 s, 1]=1.1xl06 Pas- 1 and s=O.3 are independent.
The results of the numerical solution for start-up problem (ii), F = constant, differed little from the asymptotic solution (6.34). The results of
calculations are shown in Figs 6.3-6.7 and 6.9-6.13 by lines. The solid
lines stand for the case n = 1. As seen from these figures, the theoretical
description is basically in agreement with experimental data for the
region of moderate nonlinear behaviour of the liquid. The main defect of
the theory is a limited increase in the relative viscosity, l1er = (1'/(KI1), in
steady-state extensional flow. In calculations with n= 1, l1er~6, whereas
in Fig. 6.3 this ratio is as high as 8, and in Refs 190, 274 and 281 it
reached 20. These discrepancies are eliminated in calculations when n> 1,
as shown by dashed lines in Fig. 6.3 (see also next chapter).
The relation (6.14) and coincidence of relaxation and recovery processes in time with the same initial values of stress (1' and irreversible
strain rate ep , can be derived from model (6.17) with two relaxation
modes. The limit case, (1'~311ep is also obtained from eqns (6.17) when
x~l and y~1.
The solid symbols in Figs 6.9, 6.10, 6.12 and 6.13 show the envelopes
(1'(&) and ep (&). Though for the quick adaptation approximation, the
formulae for the envelopes immediately follow from eqns (6.29), (6.46) and
(6.47), in the general case of eqns (6.16) and (6.17), the existence of the
envelopes has not been studied.

6.7 QUALITATIVE FEATURES OF SIMPLE EXTENSION AS


PREDICTED BY SOME OTHER MAXWELL-LIKE
CONSTITUTIVE EQUATIONS
In this section, we briefly analyse the predictions of Giesekus,11,12
Larson 8o and a new specification (1.56) of our gereral viscoelastic model,
in the simple elongation flow. As in the case of simple shearing, considered
in Section 5.11, only the single Maxwell models are under study here.

197

Uniaxial Uniform Extension of Polymeric Liquids

6.7.1 Giesekus Model


In simple extension, the general Maxwell-like constitutive eqns (3.1) and

(3.6) with the specifications (3.45a), take the dimensionless canonical form:
ci l dctJd,+(l- P + IJcl)(Cl -l)/cl = 2r
l
Cl dC2/d, + (1- P + PC2)(C2 -1)/c2 = - r
(0 ~ P ~ 1)
(6.48)
(no summation!)
Here the dimensionless variables are introduced as follows:

,= t/f),

(6.49)

where K is the extensional rate and all the notations are the same as in
Section 5.11. The initial conditions relating to the rest state are:
(6.50)

(i= 1,2,3)

Let us consider the steady extensional flow. It is easily proved that in


any steady flow originated from the rest state,
C2 < 1

(6.51)

cdr) and C2(r) being monotonically increased and decreased, respectively. Then, the asymptotic values of Cl (r), C2(r) and dimensional extensional viscosity, '1el(r)= O'*/r, at low and high extension rates r, are
represented as follows:
cdr)~ 1 +2r,

cdr)~2r/p,

C2(r)~ 1-r,

c2(r)~(1-fJ)r,

'1el(r)~3

'1el(r)~2/P

(r ~ 1) (6.52a)
(r~

1) (6.52b)

As seen from eqns (6.52) and an additional study, the elongation viscosity
increases monotonically when 0 < P< i and decreases at high values of r
otherwise.
6.7.2 Larson Model
In the simple extension, the general Maxwell-like constitutive eqns (3.1)

and (3.6) with the specifications (3.43) and (3.44), take the dimensionless
canonical form:
Cl l dC2/d, + B(J d(C2 -1)/c2 = - r
(6.53)
B(I d = 1 + (a/3)(cl + 2C2 - 3)
(O~a~ 1)

198

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where the same dimensionless variables (6.49) and initial conditions (6.50)
are introduced.
The steady elongation flow was considered in Ref. 80 where it was
found that independently from the value of r (0 < IX::::; 1), the elongation
stress is bounded from above by value 3/1X. This means that for high
values of r, the asymptotic behaviour of elongation viscosity is ""' r- 1
Because at small values of r, there is the Troutonian limit, the total
asymptotic behaviour of the elongation viscosity at small and high values
of r is:
(6.54)
and only one maximum on the elongation viscosity 1]el (r) exists.
The one-dimensional instability of Larson equations (6.53) against long
wave disturbances was found in Ref. 266 in the regime of relaxation after
step-wise sudden extension. The reason for this instability results from
modelling of the elastic stress-strain dependence, o"(A), because the necessary and sufficient conditions of elastic stability in this situation, do"(A)/
dA > 0, is violated for the Larson model for higher values of extension
ratios A.
6.7.3 Steady Simple Elongation for a New Constitutive Equation
Consider now the new constitutive eqn (5.85). In simple elongation, it can
be represented as follows:
A-I dAjdr +(1 2/3)n [(1 2/3)A 2 -1J/(2A 2) = r,

1]el = (A 2 - A-I )(1-1X + IX/A)/r

12 =2A-A- 2

(O::::;IX~

(6.55)
1)

where the same dimensionless variables (6.49) are introduced and the
initial condition Alt=o = 1 for the start-up extensional flow should be used.
Asymptotic behaviour of variables in steady extensional flow is of the
form:
A~1+r,1]e1~3

(r~1);

(p = 1/(n + 1),

(r~1)

q=(1-n)/(n+ 1

(6.56)

Asymptotical formulae (6.56) show that elongation viscosity monotonously increases if - 3 < n < 1 and has a decreasing branch at high
values of r when n> 1. In both cases, the elongation stress 0" is a
monotonously increasing function of r.

CHAPTER 7

On Hardening Phenomena in Flows of


Polymeric Liquids

7.1 INTRODUCTION
This chapter presents experimental evidence related to hardening and
flow slow-down phenomena in intense extensional flows of polymeric liquids, which have been preliminarily discussed in the previous
chapter. These effects are usually accompanied by very large elastic
deformations and seem to result from them. More attention is given
to extensional flows where, unlike shear flows, these effects can be
observed in 'pure form', i.e. without the interaction of polymers with
walls. Some theoretical attempts to model these phenomena in simple
extension and simple shear are also discussed. One model developed
recently is based on the assumption that a polymeric system possesses
a large relaxation time with a small elastic modulus, which appreciably affects the flow in some situations and does not in others.
Another explanation which is similar in its manifestations to the first
one and is related to an increase in the relaxation time under increasing elastic strains, is also discussed. Also considered is the possible
mechanism of distortions of potential barriers by imposed stress which
results in decreasing the relaxation time with an increase in the elastic
strain.

200

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

7.2 ADDITIONAL DATA ON EFFECTIVE EXTENSIONAL


VISCOSITY
The rheological behaviour of LDPE melts (see Table 5.1, liquid 2) under
uniaxial extension with constant strain rates K was studied in Refs 190,
274 and 275. Figure 7.1 demonstrates the data represented in Ref. 275 as
plots of effective extensional viscosity (J/K, versus time t; (J being the
elongational stress and K strain rate. In these experiments, steady extensional flow was observed for all values of K. The curve of solid circles
represents the tripled values of effective shear viscosity, 3(J 12/Y, obtained
in the linear region of deformations; (J 12 and Y being the shear stress and
shear rate, respectively. In accordance with discussions in Section 6.2. Fig.
7.1 shows that the tripled shear and elongated viscosities coincide in the
linear region of deformations, being independent of strain rate, Also, Fig.
7.1 demonstrates that at higher values of K, the ratio (J/K does not again
depend on K at the initial stages of deformation and deviates from the
linear behaviour with a further increase in time.
The main difference between the plots for LDPE shown in Fig. 7.1, and
similar ones for PIB shown in Fig. 6.3, is that for the LDPE it was
impossible to achieve steady linear elongational flow, even at very small
values of K. At the same time, the linear shear flow for LDPE was reliably
achieved with the same values of shear rate. Also, for LDPE, the steady
values of effective viscosity under extension with very small K, being all
the time slightly over 3'1, are achieved during the time i which appreciably
exceeds the equivalent time in shearing experiments. With a further
increase in K, the value of i starts to decrease, just as in the experiments
with PIB.
Also, the steady extensional viscosities for LDPE and PIB are different.
For LDPE, it initially increases and, going through a maximum, eventually decreases with growing K. No decrease in extensional viscosity was
observed in elongational flows of PIB,271 apparently due to the instrument's limitations.
Some regularities in change of the extensional viscosity with time
under K=constant, similar to those reported in Ref. 275, were also
observed in the extension of butadiene rubber and polyethylene. 192 ,276 In
the case of butadiene rubber, steady extension in the linear region was
always achieved. Also, with growing K, an increase in steady extensional
viscosity occurred, and the time of attainment of steady extensional flow
increased appreciably.192 The results of Ref. 276 are discussed in the next
section.

Hardening Phenomena in Flows of Polymeric Liquids

201

(11 'Ie

3(1f21i'10 6
( Pa[1)
..

10
..
11
II
G

'"
A

v
....-:j:""
.~oO

~.

g.. -

" V""'a
4.,.,'1'

yo
D

/'

C/ . _ _.....4_.".

A Ii,

JO

.. .

mJL-______L-______L-____~~____~______~_____
10-2

10- 1

10 3

10 0

tU)
Fig. 7.1 Experimental plots of effective viscosity (JIK versus time t for extension
of polyethylene melt (Table 6.1, liquid 2) at constant strain rate K, whose values
(S-1) are denoted by figures at corresponding curves.190.275 The plot 3(J12/Y
versus time obtained at constant shear rate y= 10 - 3 S - 1 is denoted by solid dots,
the calculated plots are shown by lines.

Thus, as established in this section, the steady extensional viscosity for


some polymers passes through a maximum when the strain rate increases.
The time of attaining the steady flow in a weakly nonlinear extensional
region considerably exceeds the similar time corresponding to the linear
region of deformations, and decreases after that with a decreasing
elongational rate.
7.3 HARDENING EFFECTS IN EXTENSIONAL FLOWS OF
POLYMERIC LIQUIDS
This section describes the effects275.276 of the suppression of elongational
flows in molten polymers by large elastic strains developing during
extension. These phenomena did not occur in extension of PIB P-20
where the irreversible strain rate, ep = d In(j&)/dt, monotonically increased in time. Below, these effects are considered under uniform
extension at a constant strain rate and a constant force.

202

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

F/ $.1Cf0'

(P",)
0.15

a)

gOO

f800

F/,fox

2700

t(,r)

6)

10-~

. . .. ..

(Pa)

0.2

I
I

I
I
0

$)

0.1

tn'

en;'
-en;

300

600

900 tid')

Fig. 7.2. The plots of engineering stress F/S o ((a)-(d), left-hand side) and the
total In (1), recoverable In Ii (2) and irreversible In (J (3) strains ((b)-(d),
right-hand side) versus time t for extension of polyethylene melt at constant
strain rates K=(a) 1.2 x 10- 3 , (b) 3.84 x 10- 3 , (c) 1.2 x 10- 2 , (d) 1 x 10- 1 S-I.

Most experiments were carried out using industrial LDPE with a


molecular weight M ~ 10 5 and within the temperature range T = 125150C. This polyethylene, being transparent above the crystallization
temperature, Tm~ 113C, becomes opaque when cooled below it. At
125C, the Newtonian viscosity for the LDPE 1] ~ 3 x 10 5 Pa s -1 and
the characteristic relaxation time e~ 100 s. This time was evaluated
as the time taken to attain steady flow in the linear region of shear
deformations with shear rate y= constant. It was also noted that preliminary long (up to 2 h) heating could slightly change the results of
extensional measurements, e.g. by decreasing the stretching force F(t)
under extension with constant strain rate K, although the dependences
F(t) remained qualitatively the same. With 15-20 min heating, no
changes in rheological behaviour of LDPE were observed. This
period of pre-heating the specimen was quite enough because the
characteristic heating time th ~ R6/a ~ 100 s; Ro = 0.25 being the initial

203

Hardening Phenomena in Flows of Polymeric Liquids

F/J'o'
10-4(Pa)
0.8

c)

o 0 0

1
I

I
I

0-2

0-3

r.z00I 0

......

7lk
1
:1 :

I
Jljlj]?
I

1111

If

Jii: 0::0

120

000

d)

d)
1

0.2

2
0.1

Jlj: j]?

I
I
I

0-2

I
I

.-$

o L-+_--,I-!-::_LI
--:S--....",.-J
5
15
25
t(.f)

1 :

15

ttj')

Fig.7.2.-contd.

radius of specimen and a = 10- 3 cm 2 s -1, the thermal diffusivity of the


polymer.
Some qualitative experiments, discussed in this section, were also
carried out with HDPE and polystyrene melts.
Let us begin with extension at a constant strain rate K. In Fig. 7.2,
the circles denote the experimental plots of engineering stress F ISo,
elastic strain In & and irreversible strain In tJ (where tJ = 8/&) versus time
t at various strain rates K. No separation of strains was conducted in
the experiment with K=1.2xlO- 3 s- 1 . The straight solid lines (1) in
Fig. 7.1 (b)-(d), show the plots of complete strain In 8 versus time t.
The time dependence of engineering stress, FISo = a{t) exp( - Kt), better

204

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

displays the solid-like properties of the melt than the true stretching
stress O"(t).
When K= 1.2 x 10 - 3 S - \ only one maximum was observed on the
dependence F(t) as in the case of PIB (see Fig. 7.4), whereas for
K?; 1.2 X 10- 3 s-1, there were already two maxima. With K=3.84 x 10- 3
and 1.2 x 10- 2 S-1 (Fig. 7.2 (b) and (c the four regions of deformation
can be introduced.
Region I starts from the beginning (t = 0) of deformation and is
characterized by a small, e.g. less than 5%, difference between the elastic,
In Ii, and total, In 8, strains. The deformation time in region I is considerably lower than the characteristic relaxation time () and the polymer is
deformed here like a cross-linked rubber; the force in this region being
increased.
In region II, where the extensional flow of polymer develops, the
irreversible strain rate ep grows and the stretching force, having passed
through the maximum, starts to decrease.
Region III begins at the inflection point on the descending part of plot
F(t), which corresponds to the inflection point on the growing part of the
dependence
In this region, the irreversible strain rate ep starts
decreasing with increasing time and approaches zero, meaning that
In ~ constant. This means, in turn, that irreversible flow in this region
is suppressed. Also in this region, the force increases again and the molten
polymer is deformed once again as a crosslinked rubber under the
condition of Ii; = constant.
In region IV, a repeated flow with ep>O develops again.
It should be noted that in all four regions, the polymer remains equally
transparent and is deformed uniformly.
When K= 0.1 S -1 (Fig. 7.2(d)), the minimum on the plot F(t) disappears
and after the first force growth, the second one starts immediately, with
region II being practically absent. Thus in this case, the initial (t < ()
region of elastic deformations during the first period, merges with the
region of repeated hardening and in both regions the irreversible strains
do not develop at all (Fig. 7.2(c.
With the temperature rising to up to 150 DC, the character of dependences
F(t) does not change and the curves in the first region are reduced to one
master curve using the time-temperature in variance with an activation
energy of viscous flow E ~ 65.4 kJ mole - 1. In the region of the second maximum, such a reduction results in a certain divergence between the curves.
Figure 7.3 shows the plots of effective elongational viscosity, 1]e1 = 0"IK
(symbols 1-7), and tripled shear viscosity, 31] = 30" 12ly (symbols 8-10)

Pet).

Hardening Phenomena in Flows of Polymeric Liquids

." "

0..

"*

..0

-1

-2
- 3
- 4-5

205

.;. -6
~ -7

-8
-9

A - fO

Fig. 7.3. The plots of effective viscosity, (JIK in extension and 3(J12/Y in simple
shear versus time t for deformations of polyethylene with constant strain rates K
and y. Symbols 1-7 denote the values of K: 1,2 x 10-1; 2,1 x 10-\ 3, 5.65 x 10- 2 ;
4, 3.84x 10- 2 ; 5, 1.2 x 10- 2 ; 6, 1.42 x 10- 3 ; 7, 1.2 x 1O- 3 s- 1. Symbols 8-10
denote the values of y: 8, 1.42 x 10- 3; 9, 1.13 x 10- 1; 10, 3.56 x 10- 1 s-1.

versus time t for LDPE at various strain rates and at temperature


T = 125C. Here (J and (J 12 are extensional and shear stresses, and K and
yare extensional and shear strain rates. The solid lines denote the results
of calculations discussed in the next section.
When K ~ 1.2 X 10- 3 s-1, extensional viscosity l'/e1 increases monotonically with time in the regions of both force growth and decline; in the
region of force growth, both l'/e1 and (J increase exponentially as exp(Kt).
When K < 1.2 x 10- 3 S - 1, it was not possible to make reliable measurements of stretching force.
Symbols 8-10 in Fig. 7.3 show the typical behaviour of effective
viscosities in the simple shear discussed in Chapter 5: the appearance of
maxima and the decrease in time of steady-state attainment with a
growing shear rate. Symbol 8 shows the experimental curve, which does
not depend on shear rate and corresponds to the data obtained from the
linear behaviour of the melt. In accordance with the results of the

206

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

previous section, this curve should coincide with the curve for '101 in the
linear region of deformation.
Symbol 7 shows that even in the case of minimum reliable extensional
rate, K= 1.2 x 10- 3 s-1, the LDPE melt demonstrates nonlinear behaviour at higher times, attaining the steady flow at a time which considerably exceeds the corresponding time in the linear region. It should be
reminded that for PIB, this time decreases with growing K, beginning
from the region of linear deformations.
Two maxima on the dependence F(t) were also observed in the
extension of the HDPE melt where, after secondary growth of F(t), the
sample was ruptured with a neck propagation. A repeated growth of the
force immediately following the first one, similar to that shown in Fig.
7.2(d), was also observed in the extension of the polystyrene melt, i.e.
non-crystallizing polymer. Though for these two polymers, the recovery
strains were not measured, if the hardening phenomena can be judged by
the repeated increase in force, the evidence provided here shows that this
effect is very typical for polymeric liquids of various nature.
Thus, following the secondary hardening either a secondary flow
develops or melt rupture occurs. This rupture, in principle, may also
occur in the regions of deformation times which are below the relaxation
time () and therefore rupture may not characterize the region of secondary hardening. Further discussion of the rupture of polymer melts under
extension are given in Chapter 12.
It is of interest to note that for the extension of PIB at room
temperature, discussed in Chapter 6, the hardening effects were not
observed in the whole region of elongational rates under study. In this
region, the dependences F(t) for PIB had only one maximum, exactly as
for the LDPE at small strain rates, and then monotonically decreased
(Fig. 7.4(a)), with the irreversible strain rate e p = dIn P/dt monotonically
increasing (Fig. 7.4(b)). But it is still possible that the hardening phenomena for PIB could be observed at higher elongational rates.
Although the extensional plots presented in Figs 7.1 and 7.3 look quite
similar, unlike the data reported in Refs 275-277, the effects in the
pre-stationary region of extension discussed in this section were, obviously, less pronounced.
Now let us consider the hardening effects in extensional experiments
under a given constant force. 282 The experiments were carried out with
the LDPE at 125C and PIB P-20 at 44C whose relaxation behaviour
was studied at 22 C (see Section 6.3.3). It is worth mentioning that
at such temperatures, these polymers have approximately the same

207

Hardening Phenomena in Flows of Polymeric Liquids

1'/$0"8
(0-

(a)

(Pa)4

(nl,e", tX

tn;

Cl

Cl

Cl

Cl

0
0

2,0
Cl
Cl

"

20

"0

80

60

ttf)

30

60

90

'b

120

ted

Fig. 7.4. The plots of (a) engineering stress and (b) components of strain:
- - - , total, In 8; 0, recovery, In Ii; and ., irreversible, In 13, versus time t for the
extension of PIB P-20 at 22C.

Newtonian viscosity, 1'/ ~ 3 X 10 5 Pa s -1, and relaxation time f) ~ 100s, as


were determined in shear experiments. Also, the flow curves for both the
polymers in the studied region of shear rates, y< 1 s - 1, did not differ
greatly (Fig. 7.5).
The results of experiments for the polymers in simple elongation are
shown in Fig. 7.6. These are the plots In 8, In Ii and In /l versus time t for

fO~
0

tJ.

Ii
Fig. 7.5.

10-2

fO- 1

tOO j(rl)

Flow curves: for LDPE at 125C (.6) and for PIB at 44C (0).

208

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

(a)

,,~

(b)

<~.

~
" !.A;)'

<~

<~,2

<"Q'

-I

'10

80

120

16'0

t(6)

Fig. 7.6. Experimental plots of total In e, recoverable In Ii and irreversible In p


(dots 1,2,3 respectively) versus time t in extension at constant forces: (a) LDPE,
(b) PIB. Dots 1', 2', 3' denote initial stresses rIo = 1.0 X 104 Pa, 1.5 x 104 Pa and
2.0 x 104 Pa, respectively.

LDPE with 0'0 =(1.0, 1.5,2.0) x 104 Pa shown in Fig. 7.6(a) and PIB with
= 1.0 X 104 Pa in Fig. 7.6(b). Here 0'0 =F/S o; F being the given constant
force and So the area of the initial cross-section of the specimen.
In spite of the proximity of the flow curves, as well as the basic
parameters e and 1'/, the rheological behaviour of these polymers in simple
extension is quite different. The time taken to attain the maximum
possible strain in these experiments, In em ~2.8, under 0'0 = 1.0 X 104 Pa is
six times higher for the LDPE than that for the PIB. The dependence
In e(t) for PIB results in an increasing strain rate, K(t) = din /dt, while for
LDPE, K(t) decreases monotonically within a large region of e and
becomes approximately constant at large times. With a further increase in
time, a growth of K(t) is observed for both LDPE and PIB.
Figure 7.6 also demonstrates a qualitative difference in plots ofln &and
In versus time for these two polymers. Especially remarkable is the
existence of a plateau for LDPE where In(e/&) ~ constant, i.e. the flow of
the polymer is almost not developed. On this plateau, the polymer
deforms almost like a solid elastic body. The height of the plateau
remains approximately the same for all three values of 0'0' This hardening
0'0

Hardening Phenomena in Flows of Polymeric Liquids

209

process results in the growth of the effective viscosity, (J/K, with the
development of subsequent flow, where the development of irreversible
strain In fJ leads to an increase in the strain rate K. Under (J 0 = 104 Pa, at
the maximum total strain In 8m ~ 2.8 achieved in the experiments, the
elastic component of strain, In (l(m, for PIB is 1.7 times smaller than that
for LDPE. This is obviously due to the hardening in the LDPE melt
under extension.
It should also be noted that at the initial instant of loading, a very
sharp growth of deformation occurs for both polymers, which is approximately equal to 10% of the initial length of the specimens.
Figure 7.7 displays the plot In eX versus In 8 for the LDPE melt, which,
as in the case of PIB (Fig. 6.8), does not depend on the initial stress (J 0 but
is qualitatively different from the latter. The extension of purely elastic
solids is shown on the figure by the straight line with a 45 0 slope.

0-1
D.-2
f

0-3

Fig. 7.7. Experimental plots of elastic recovery In IX versus total strain


In obtained in extension of LDPE at various constant forces. Symbols 1, 2
and 3 denote initial stresses (j 0 = 1.0 X 104 Pa, 1.5 x 104 Pa and 2.0 x 104 Pa,
respectively.

The extension of LDPE is accompanied by an accumulation of


noticeable irreversible changes which were detected in Ref. 276 by a
comparison of the results of extensional experiments for fresh and
repeatedly deformed samples with an identical thermal pre-history. In
spite of the quantitative changes in the extensional dependences for the
repeated experiments, the qualitative features of the experimental results

210

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

remained the same. Unfortunately, the experimental results obtained in


Ref. 276 did not reveal the effect of irreversible changes in a polymer on
the repeated mechanism of flow.

7.4 MODELLING THE HARDENING PHENOMENA


IN EXTENSIONAL FLOWS OF POLYMERIC LIQUIDS

7.4.1 General Remarks


The brief analysis of experimental data presented in Sections 7.2 and 7.3
points to the existence of two physical mechanisms with opposite
manifestations, which dominate in high Deborah extensional flows of
liquid polymers. One of them promotes the hardening phenomena and
slows down the extensional flow, while the other accelerates it.
The first mechanism, related to the hardening phenomena, manifests
itself at K= constant in such effects as: an appreciable increase in the time
of attainment of a steady flow in the nonlinear region of deformation as
compared to the linear one; the appearance of the second maximum on
the time dependence of stretching force F(t); and a decrease in the
irreversible strain rate ep during extension. In extension under F = constant, there exists a remarkable difference in the time attainment of a fixed
deformation for various polymers whose flow curves in shearing are
practically identical.
The second mechanism manifests itself in the passing of an effective
steady elongational viscosity, through a maximum. In extension where
the steady effective viscosity can considerably increase, this mechanism
acts in the opposite direction to the viscosity increase and therefore is
easily detected. In shearing, where the viscosity decreases, this mechanism
may be almost unnoticeable. It should be noted that this mechanism may
be associated not only with the reversible structural changes in flowing
polymers, but also with effects of mechanodegradation,288,289 i.e. with
rupturing polymeric chains. The results of repeated extensions 276 discussed at the end of the previous section can serve as indirect evidence of
the mechanodegradation.
Two attempts at the modelling of hardening phenomena have been
proposed in the literature; both being phenomenological and, to some
extent, speculative. One of them, proposed in Ref. 8, assumes that a
characteristic relaxation time of a liquid can exponentially grow due to a
decrease in the activation entropy because of orientation phenomena.
This approach is described by eqn (1.58). Another model, proposed

Hardening Phenomena in Flows of Polymeric Liquids

211

recently,300 introduces a special relaxation mode with a very low elastic


modulus and a very high relaxation time. Both of them, being rough
enough in descriptive abilities, are considered below.
In an attempt to model the decrease of elongational viscosity,
Euring's idea about the distortion of potential barriers by stress was
employed as described by eqn (1.59). It should be noted that this
reversible mechanism is very different from mechanodegradation which
cannot be described by such an approach. To the authors' knowledge, no
distinct attempts were made in the literature to separate these effects.
7.4.2 Modelling of Hardening as the Relaxation Transition
As assumed in Ref. 8, the characteristic processes of relaxation and
viscous flow in concentrated polymeric liquids are mainly determined by
the activated formation and break of secondary linkages between the
neighbouring parts of macromolecules. Thus, the characteristic relaxation
time can be expressed as: 62
(J* = (Jo(T) exp(dJ*/RT),

(7.1)

dJ*=dU*-TdS*

where dJ*, dU* and dS* are the specific activation free energy, internal
energy and entropy, respectively; Po is the density in the undeformed state
and T is the temperature (Kelvin). It is assumed that in strong flows, the
term dS* equal to the order of entropy of the segment participating in
the transfer act, diminishes as a result of the orientation effect of the
mechanical field. Therefore it was approximately assumed that
dS* ~ - f3w(I 1> 12 ),

f3~M */Me(O~f3 ~

w~2W.(1 1>

12 )//1

1),
(7.2)

where Ws = [W(I 1> 12 ) + W(I2, 1 1 )]/2 is the 'symmetrized' elastic strain


energy function, /1 is the Hookean modulus, f3 is the flexibility parameter
estimated independently, M * is the molecular weight of a statistical
segment characterizing the thermodynamic flexibility of the macromolecular chains, and Me is the molecular weight of the polymer chain
between two entanglements. The use of a symmetrized potential is, of
course, not necessary, and was introduced in Ref. 8 for the reason of
potentiality of irreversible forces. The values of M * and Me are estimated
as follows:
M*=vm,

Me = Mcr/2

(7.3)

212

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where m is the molecular weight of a monomer unit, v'" 5-10 is the


averaged number of monomer units in statistical (Kuhn's) segment and
Mer is the critical molecular weight when the exponent index in the
dependence of viscosity on molecular weight has a step-wise increase.
Also, when using Euring's concept of the distortions of potential barriers
which govern the activation process, then, say, in the case of simple
stretching with stress a, we have:
(7.4)
Combining formulae (7.1), (7.2) and (7.4) yields:
()* = ()(T) exp[pw(I b 12) - fa J,

()(T) = ()o(T) exp(~U6/RT)

(7.5)

The qualitative behaviour of a single Maxwell model with an account


of hardening phenomena was examined in Ref. 290 in both simple shear
and simple extension. We consider here only the hardening effects in
simple extension. The effects in simple shearing will be discussed in
Section 7.5.
The constitutive equations for the simple extension of a single Maxwell
model with neo-Hookean elastic potential take the form:
d2+(2+1)(2 3 -1)
dT 62 exp[pF(2)J

AT

(7.6)

where
r=K(),

T=t/(),

a*=a/p,

F(2)=(22 +22+22 -1 +2 -2 -6)/2

(7.7)

where G is the elastic modulus, and () is the relaxation time. The


steady-state curves a*(r) are represented in Fig. 7.8(a) for three values of
parameter p; the upper branches of these curves, shown by dashed lines,
are unstable. For any particular value of p from the interval (0, 1J, there is
only a finite region of r (or 2): r~r~(p) (2~2*(P)) where steady
extensional flow exists; the critical values of r~(p) and 2*(P) are decreasing functions of p. When r > r~(p), r = constant and T~ 1, eqns (7.6)
result in the asymptotic formulae:
2~exp(rT),

(7.8)

which corresponds to the pure elastic regime of deformation. This is a


simple manifestation of the 'fluidity loss' concept proposed for explanation
of the hardening phenomena in Refs 291-294. According to eqns (7.6), this

Hardening Phenomena in Flows of Polymeric Liquids

80
40

(a)

" ",-

213

~=0'1

-..;

. . . . ~=O8 . . . . ........

~=1-o-- - - . . .
r
OC'~-~~0~'5~==~1'0~---1'~5~

.-o
(b)

(c)
Fig. 7.8. Hardening phenomena as an relaxation transition for a single Maxwell
model: (a) dimensionless extensional 'flow curves' for various values of p; (b)
(P=O.l) and (c) (P=l) are dimensionless start-up flows (right branches) and
relaxation curves (left branches) for various constant extensional rates.

is a relaxation transition from liquid to solid behaviour due to the sharp


increase in the relaxation time induced by orientation phenomena.
The transient behaviour corresponding to start-up extensional flow
and stress relaxation are shown in Figs 7.8(b) (P=O.1) and 7.8(c) (P= 1). It
is of interest to note that when P= 1, eqns (7.6) predict practically linear
behaviour of elastic liquids up to r=r~ which has been observed in the
extension of polymer melts with narrow MWD.280 The same feature
occurs in simple shear.
According to the concept of fluidity loss, at r > q(P) rapid stress
growth should result either in mechanodegradation of the polymer melt
with a following decrease in its relaxation time and the occurrence of a
secondary flow, or in the rupture of a stretched specimen by the
mechanism characteristic of the cross-linked rubbers.
The experiments 38 ,275,276 considered in the previous section, were
semi-quantitatively described in Refs 38, 285 and 295 using the approach
of relaxation transition extended for the multimodal Maxwellian model
as developed in Ref. 8.

214

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

A more detailed comparison of the literature data and the predictions


of the multimodal approach with allowance for hardening phenomena,
was made in Ref. 296. The following problems were considered: the stress
growth at a constant elongation rate or at a constant pulling speed, the
elastic recovery under elongation flow, and the transient tensile stress
behaviour in multi-step loading-relaxation experiments. The calculations
were compared with data for LOPE melt UPAC A at 150C,274 PIB
P-20 at 22 C,286 PP melt at 200C,297 PS Shell TC 3-3- at 160C 298 and
PIB Oppanol B, 50 at 23C. 299 The remarkable feature of Ref. 296 was
that the relaxation parameters of the model (relaxation times and elastic
moduli) were found in the linear region of shear or extensional deformations, whereas the only parameter {3 which characterized a strongly
nonlinear hardening effect, was calculated according to formulae (7.2) and
(7.3) from known properties of the polymers. Reference 296 found 'the
resulting predictions to be in fair agreement with the available experimental data for different polymer melts', which testifies in favour of the
description of hardening phenomena by the relaxation transition. The
slight disagreement with data at extremely high elastic strains, accumulated in strong extensional flows of polymer melts, may be related to the
secondary acceleration of flow due to the mechanodegradation or the
distortion of potential barriers by stress, which were not taken into
account in calculations. 296
7.4.3 Modelling of Hardening Phenomena with a Long Time and
Low Elastic Modulus Relaxation Mode
This mechanism was first predicted by a molecular theory133,134 for
monodisperse polymers, and recently was proposed in Ref. 300 for the
description of the hardening phenomena. Along with the inequality (1.69)
proposed for the multimodal approach, it was assumed that in eqns
(6.16), one (say, the first) relaxation mode has a large relaxation time 8 1
and low elastic modulus I-lb and
n

111 ~

k=2

11k

(7.9)

where we assume, as usual, that the inequalities 11k ";311k+1 hold for k> 1.
If, in the course of deformation, the first relaxation mechanism begins
to dominate in stress over the others, the effective relaxation time of the
whole system increases. Then in the time interval !1t of the domination,
the effective Deborah number Del ;:::;8d!1t ~ 1 and the system is almost

Hardening Phenomena in Flows of Polymeric Liquids

215

deformed as an elastic solid. Due to ineqn (7.9), the first mechanism


contributes very little stress to the linear region and in this case, the
characteristic relaxation time of the polymer can be taken as ()2 (()2 ~ ()1)'
It is convenient to begin applying this idea to a qualitative consideration of extension in the viscoelastic system at such a region of constant
strain rates K that r = ()2K ~ 1. Because in the extension, the growth of
viscosity occurs independently in each relaxation mode, there can exist
such an interval of strain rates where the effective viscosity increases only
due to the deformation in the first relaxation mode. In other relaxation
modes, the effective viscosity does not change since they are in the linear
region of behaviour (()k K ~ 1,
k> 1). Thus, the effective viscosity in the
first relaxation mode may exceed the sum of viscosities, k~ll1k while the
flow will develop more slowly than that predicted by the mode with
relaxation time ()2' In simple shear where the effective viscosity in the first,
long-term, mode decreases in a similar situation, this mode contributes
almost nothing to the effective viscosity.
Based on this approach, results are presented below for the numerical
modelling30o.301 of extensional flows of polymers with hardening phenomena. Both the regimes of extension, K= constant and F = constant, are
considered and compared with the experimental data presented in
Section 7.3. The calculations were based on eqns (6.16) with relaxation
times ()~ described by eqn (1.59b) and using a two-mode model with a
viscous element. The following parameters were used in the modelling:
()2 = 100 s,

n2 = 1.1,
(l1k=/1k()k,

/11 = 8.8 Pa,

Y1=2xlO- s Pa-1,
111=0.7x10 s,

/12 = 4.1 0 3 Pa,

Y2=0,

112=4x10 s,

s=7.8xlO- 2 (7.10)
11=5.1x10 s Pas- 1)

In calculations of start-up extensional flow with K= constant, the initial


conditions (6.19) were employed, while with F = constant, the initial
conditions (6.25)-(6.26) were used. The results of the calculations are
represented in Figs 7.2, 7.3, 7.7, 7.9 and 7.10 by solid and dashed lines.
In Fig. 7.2, the curves denote the calculated plots of stretching force F,
recovery strain In eX and irreversible strain In fJ versus time t, with various
values of constant extension rate K. The first maximum on the dependence F(t) is due to the relaxation phenomena in the mode with parameters
()2 and /12, while the second maximum manifests the first relaxation mode
with the large relaxation time ()1, small elastic modulus /11 and viscosity
111' The start of the flow slow-down, with an increase in the irreversible
strain rate ep = dIn fJ/dt, corresponds to the beginning dominance of the

216

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Fig. 7.9. Calculated plots of effective extensional viscosity versus time with
allowance for both the effects of long time relaxation mode and distortions of
potential barriers (solid lines) and without these effects (dashed lines). Curves 1-8
correspond to the strain rates K: 2 x 10 -1, 1 x 10 - \ 5.65 x 10 - 2, 3.84 x 10 - 2,
1.2 x 10- 2 , 3.84 x 10- 3 , 1.2 x 10- 3 and 3.84x 1O- 4 s-\ respectively.

if;

0.8

80

120

Fig. 7.10. Calculated (lines) and experimental (points) plots of elastic recovery
In & versus time t for extension at constant force with (To = (2.0, 1.5, 1.0,
0.5) x 104 Pa. The symbols correspond with the data in Figs 7.1-7.4.

Hardening Phenomena in Flows of Polymeric Liquids

217

mode with the larger relaxation time 8 1 . It is worth noticing that the
mechanism of distortion of potential barriers by stress does not qualitatively affect the results shown in Fig. 7.2. Also, the second maximum on
the dependences F(t) was obtained in Ref. 300 with the assumption
n1=n2=n.
Figure 7.3 shows the results of calculations of effective elongation ai,
U/K, and tripled shear, 3U12/";, viscosities depending on time t, at various
extensional rates K and shear rates ";. Here the manifestation of the mode
with the larger relaxation time 8 1 and low elastic modulus, explains why
the time of attainment of the steady flow in extension with, e.g.
K=1.2x10- 3 S- 1 is much higher than that for the linear region of
deformation in simple shear with the same strain rate, "; = K. In the linear
region of deformation, this mode is unnoticeable and hardly affects the
effective viscosity in shearing.
The delay in attainment of a steady nonlinear elongation flow and its
following decrease with K growing, was also confirmed in calculations
with nk = nand Yk = O. The case Yk;6 0 is considered here only to describe
the possible decreasing branch of steady extensional viscosity. Though no
quantitative comparison has been carried out with the data 275 shown in
Fig. 7.1, they are also in complete qualitative agreement with this
approach.
The results of calculations shown in Fig. 7.9 demonstrate the importance of both the large relaxation mode and the mechanism of decreasing
relaxation time. The model calculations, including both the effects, are
shown in Fig. 7.9 by solid lines, and those without the effects, by dashed
lines. Unlike the experimental data shown in Fig. 6.3, the steady elongational viscosity calculated without the second effect, increases unlimitedly, and when a large relaxation time mode is taken into account, it
practically dominates. In the second relaxation mode the effect of
distortion of potential barriers by the stress is insignificant throughout
the region of experimental data.
Figure 7.10 shows a comparison between the results of calculations
(lines) of recoverable strain In Ii and experimental data (symbols, see Fig.
7.6) for various initial stresses Uo = F/S o in extensional deformations at
constant force F; So being the area of the initial cross-section of the
specimens. The inconsistencies between the data and calculations at low
values of Uo can be attributed to a certain change in the rheological
properties of the large relaxation time mode that might occur during a
1-year break between the experiments with K=constant 276 and F=constant. 282 The results of these simulations showed that the higher are the

218

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

parameters nk and 0 1 at fixed values of 0"0 and 0 3 , the slower is the


process of extension.
The results of calculations of the dependences In e(t) are very similar to
those for In &(t). It should also be noted that a sharp initial growth of the
total strain In e mentioned above when the data of Fig. 7.7 were
discussed, is apparently due to the occurrence of the inflection point at
the dependence In e(t) when nk> 1 (see Fig. 6.16).
7.5 MODELLING OF HARDENING PHENOMENA IN SIMPLE
SHEARING OF ELASTIC LIQUIDS
In shearing flows, the hardening of polymeric liquids can result in
more complicated phenomena associated with wall slip which, in turn,
leads to a complete change in the character of flow. The wall slip
phenomena is discussed in Chapter 12, this section being confined only
to the flows with the stick conditions at the wall. The same two
mechanisms of hardening as proposed in the previous section are
considered below.
7.5.1 Hardening as a Relaxation Transition

In order to clarify qualitatively the hardening phenomena which could


occur in simple shearing, let us consider once again a single relaxation
Maxwell mode under given shear rate y290 (a brief analysis of the
problem is also considered in Ref. 296 ) The dimensionless equations take
the form:
(O~P~

1)

(7.11)

where
W=Cll

+C22-2
0"2=C22-

(7.12)
1

In eqns (7.11) and (7.12) the stresses are scaled by elastic modulus 11, and
time and shear rate, by relaxation time O.

219

Hardening Phenomena in Flows of Polymeric Liquids

Equations (7.10) and (7.11) have the steady solution:


xZ+ 1
-- -J--;==z'
ell

1-x

(7.13)

The dimensionless flow curve, u? z(r) and the first normal stress difference, u?(r), are shown in Fig. 7.11 for three values of {3. When {3 > 0, the
flow curve is double-valued; the upper branch being linearly unstable.
Thus a steady flow exists only within a bounded region, 0:;:; r O < r m({3),
where r m({3) is the maximum of rO(x) with respect to x.
\
1.0""---------

JO.5

0-

(al

10

,
I

\3
\

0.5

1.0

1.5

(bl

Fig. 7.11. Steady-state theoretical plots (a) a?2(r) and (b) a?(r) according to
eqns (7.13); for some values of parameter p: curve 1,0.0; curve 2, 0.1; curve 3, 1.0.

When r>r m(f3), steady flow does not exist and if the stick conditions
at the wall hold, the stresses grow unboundedly with asymptotic values:
(7.14)

which correspond to the pure elastic regime of deformations. This


asymptotic behaviour was treated in Ref. 290 as a relaxation transition
from flow to the high elastic behaviour of polymeric liquids after their
fluidity loss.
The numerical solution of the problem of start-up flow z90 showed,
however, early instabilities of eqns (7.11) with critical values
r *({3) < r m({3). Figure 7.12 demonstrates that the difference between r *({3)

220

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

10

LE
L*

b\
\

,,
"-

"-

"
0.5

1.0

Fig. 7.12. Dependences of critical shear rate on parameter [3 in simple shear


flows; - - - , critical shear rate 1 m([3) obtained from linear stability analysis;
- - -, critical shear rate 1*([3) obtained in nonlinear numerical studies.

and r m(f3) at small values of f3 is high enough. This difference was


explained in Ref. 290 by qualitative analysis of the global phase diagram
in the phase space {c 11 , C 12}' It should be also noted that the critical
values r * in simple shearing exceed (sometimes considerably) the critical
values of r~ for the simple extension, because the extensional flow is
accompanied by stronger orientation of polymer than the shearing one.
Some results of numerical calculation of the set (7.11), (7.12) for
start-up flow and relaxation are shown in Figure 7.13. Recent multimodal

10
1:

Cal

15

1:

Cbl

Fig. 7.13. Calculated transient dimensionless dependences for (a) shear stress
adt) ([3=0.1, 1*=0.918), and (b) first normal stress difference adt) ([3=1,
1 * =0.618) for some constant values of dimensionless shear rate 1: curve 1,
1 = 1 */2; curve 2, 1 = 1 *; curve 3, 1 = 1.2 1*. The left-hand side of the curves
correspond to start-up flow, the right to relaxation.

Hardening Phenomena in Flows of Polymeric Liquids

221

analysis 302 of the critical regimes of deformation of elastic liquids in


shearing flow confirmed the above results of the critical values r m(f3) and
r *({3) represented in Fig. 7.12.
More discussions of this modelling of hardening phenomena in die
flows of viscoelastic liquids is given in Chapter 8.
7.5.2 Modelling of Hardening Phenomena with a Long Time and Low
Elastic Modulus Relaxation Mode 40
As noted above, this section initally omits the effect of stress distortion
of potential barriers because it contributes almost nothing to shear
viscosity.
It is easy to examine the influence of this mode in steady shearing, since
if it is negligible in a steady situation, the same is true for transient flow.
To simplify the calculations we consider further that 0* = 0 = constant,
nk=n and N=2. Then using the asymptotic formulae (5.75), we can
represent the dimensional effective viscosity ilk = (J 12,k/Y in the kth relaxation mode as follows:

Since, due to ineqn (5.76), 1 ;:::;n;:::;2, the contribution of the long-term


relaxation mode in the effective viscosity decreases quickly with Y growing.
For the same reason, the long-term relaxation mode hardly affects the
dynamic modulus G"(w) in the intermediate region of frequencies w.
Most important, however, is the behaviour of the first normal stress
differences, (Jl,b in the relaxation modes. Consider first the case when
yOI ~ 1. Then the first normal difference is represented as:
(7.16)
and due to the inequality IhOl~1J202' the large-scale relaxation mode
dominates at the small shear rate. It also continues to dominate in the
(very large) intermediate region of shear rates where 0 1 y~ 1 and O2 y~ 1.
In the region of high shear rates, when O2 y~ 1, the asymptotic formula
(5.75) for (Jl yields:

(=111 [01Jn12
112

O2

(y02 ~ 1)

(7.17)

Thus if (~1, the contribution of the long relaxation mode in the first
normal stress difference at y02 ~ 1 is insignificant; if, however, (> 1, its
contribution is essential.

222

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In the case when n1 #n2, y02 ~ 1, and the stress distortions of the
potential barriers are taken into account, the asymptotic formula (7.17)
for the first normal stress difference changes to the following:
(J 1 ~

L (,uklnk)(4yOSk/ 2 exp[ -

(Yklnk)(4yOk)"k/2] ==

L (J 1,k

(7.18)

where (J1,k are the contributions in (J1 from the kth relaxation mode. Let
us estimate the contribution of the long-time relaxation mode in (J 1 for
the real case of shearing polyethylene when the parameters are given by
eqns (7.1 0) for y = 0.1 s - 1 (yO 2 = 10). Then, if Yk are neglected,
(J1,1 ~ 1.1 X 10 5 Pa and (J1,2 ~ 6.9 X 10 3 Pa; if Yk are taken from eqns (7.11),
(Jl,l~4xl04Pa and (J1,2~6.9x103Pa. These results show that the
long-time relaxation mode for the polyethylene contributes overwhelmingly to the normal stress in all regions of shear deformation and can be
easily detected by measurements of normal stresses. This relaxation mode
can also be revealed with the help of dynamic tests at W01 ~ 1 where
according to the Cox and Merz rule,231 G'(w)~(Jd2Iy=w' Until now,
these contradictions were not resolved.
The attempts to reveal the large-scale relaxation mechanism in simple
shear with small dimensionless frequencies and shear rates are fairly
complex even if 0 1 ~ 104 s, though available evidence (see, e.g. Ref. 581)
suggests that such relaxation times can be measured in dynamic experiments. Figure 7.14 shows the plots G'(w) and G"(w) for linear hydrogenated polybutadiene. 581 According to eqns (5.21), the asymptotic
relations
dlnG"/dlnw~l,

d In Gild In w~2

(w~O)

(7.19)

correspond to the slopes of the curves equal to 45 and 60, respectively.


Figure 7.14 shows, however, that if for Gil the slope is equal to 45 for any
w < 3 x 10 - 1 S -1, the slope for G' is equal to 60 only within the range
3 x 10 - 2 ~ W ~ 3 x 10 - 1 S - 1 and then decreases again with decreasing w.
This fact suggests that there can be a long-term relaxation mode with a
very low modulus and relaxation time 1 ", 10 3 s.
Besides, the maximum on the time dependence of the first normal
stress difference (J 1(t) along with the long-time approach of (J 1 to steady
state, shown in Fig. 7.15,303 may also be attributed to a presence in the
system of a long-time relaxation mode with a very small elastic modulus.
At the same time, Fig. 7.3 shows that no deviations from the normal
behaviour on the time dependence of shear stress (J12(t) have been found.
It should be noted, however, that the results of transient normal stress

223

Hardening Phenomena in Flows of Polymeric Liquids

C OOCO

0&'8 000
o8

Fig. 7.14. Plots of dynamic moduli G' and G versus frequency of oscillations w
for linear hydrogenated polybutadiene, M ~2 x 10 5 : 581 0, G'; D, G
U

r;,z.
~1

(Pa')

JO

20

fO

0.2

0.4

0.6

0.8

1.0

t(.r)

Fig. 7.15. Plots of shear stress 0"12 (curve 1) and first normal stress difference
(curve 2) versus time in shearing with constant shear rate y= 111 S -1 for 0.2%
aqueous solution of a biopolymer (Kelsan),303 10 min after agitation of the
solution.

measurements in very viscous liquids, obtained in cone-plate rheometers


(as in Ref. 303), should be considered very carefully because of the large
characteristic time of the measuring device (see Section 4.2.5).

CHAPTER 8

Flows of Polymeric Viscoelastic Liquids


in Channels and Pipes

8.1 INTRODUCTION
Flows of polymeric viscoelastic liquids are of particular importance in
any type of polymer processing, especially in extrusion. They are also of
interest for obtaining reliable data from capillary rheometry, as is pointed
out in Chapter 4. It is important to mention that controversy exists
between theoretical predictions of rectilinear flows and a few experimental studies of steady viscoelastic Poiseuille flows in long tubes with
arbitrary cross-sections. This is the subject for discussion of Section 8.2.
Section 8.3 considers the steady and unsteady Poiseuille flows in long
pipes with the simplest cross-sectional geometries: circular pipes and die
slits. Also, some possible hardening phenomena are briefly discussed
there. Section 8.4 deals with entrance and exit die flows, where experimental data along with numerical simulations are under discussion.
8.2 GENERAL RESULTS FOR RECTILINEAR STEADY FLOWS
OF VISCOELASTIC LIQUIDS IN LONG TUBES
WITH ARBITRARY CROSS-SECTIONS
Ericksen 304 was the first to discover the curious fact that the rectilinear steady flow of non-Newtonian liquids is generally impossible

Flows of Viscoelastic Liquids in Channels and Pipes

225

(see also Refs 305 and 306). For general viscoelastic liquids, this problem
was analysed by Noll and is represented in Ref. 93 (Section 117), however, with a minor confusion of the first and second normal stress
differences. The correct results are presented in Ref. 4 (Sections 3.5
and 4.11.3). The derivation given in Refs 4 and 93 is briefly reproduced
below.
Let us assume that the rectilinear flow of an elastic liquid does exist.
In the Cartesian coordinate system shown in Fig. 8.1, the components of
velocity are:
(8.1)
Here, in accordance with the rules established in Section 4.2 for viscometric flows, the axis Xl is directed along the flow. The axes X2 and X3,

Fig. 8.1. Cartesian and viscometric coordinate systems


for general rectilinear pipe
flow of viscoelastic liquids.

being arbitrarily disposed in the cross-section orthogonal to Xb have


nothing in common with the viscometric axes defined in the simple shear.
Despite the difference between this kind of possible flow and simple
shearing, it is still possible to establish in the plane X2, X3 such
a curvilinear orthogonal system in which the problem will be locally
reduced to the simple shear flow. Namely, following Ref. 93, we introduce
an orthogonal coordinate system in the plane X2, X3 such that one
coordinate family is defined as Xl = U(X2' X3) and another, X3, is orthogonal to Xl' This new coordinate system, X'l =Xb Xl,X3, is locally related
to the simple shear flow with the basic stresses defined by matrix (4.3).
Then, using the common transformation rules and stress equilibrium

226

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

equation, one can obtain the following two equations for one velocity
distribution v(xz, X3) searched for:
V [l/(Y)Vu] = q

(8.2)

V [(O"z(Y)/yZ)Vu] = f(u)

(8.3)

where y = IVul is the effective shear rate, l/(Y) is the effective shear
viscosity, q = - OP/OXl is a constant longitudinal pressure gradient, O"z(y)
is the second normal stress difference, and f(u) is an arbitrary function of
the velocity u(xz, X3), which can also be equal to zero. As in simple shear
flow, the first normal stress difference O"dY) does not have any effect on
this type of flow. It should be also noted that in the original works (see
e.g. Ref. 93) the magnitude 0" 1 (y) was confusedly put into eqn (8.3) instead
of the correct quantity O"z(y),
For polymer melts and concentrated polymer solutions, the experimental values for the ratio of the second to first normal stress differences
decrease with increasing y and usually lie in interval -0,05 to -0,3 (see,
e.g. Refs 3 and 6). The values of O"z(y) are accurately detected by the
birefringence method, 3 and their absolute values increase with increasing
y. These results show that in a steady flow of viscoelastic liquids in long
tubes of arbitrary cross-sections, when De is high enough, it is impossible
to satisfy simultaneously both eqns (8.3) and (8.4), i.e. the rectilinear flow
does not generally exist.

There are two exceptional cases, however.


(i) if the slow flow (y--+O) in a tube is considered, for which
l/~'1=constant and O"z(Y)/yZ~'I'z(O)=constant, eqns (8.3) and
(8.4) are satisfied simultaneously; eqn (8.4) being satisfied if the
function f(u) is chosen as: f(u) = q'l'z(O)I'1. This is the so-called
second-order approximation in slow flows of elastic liquids. 6 ,93
(ii) The second case when the rectilinear flow exists, theoretically at
any values of De, are the cases of capillary, circular die and planar
Poiseuille flows for which (and only for which) y=y(U).304,307
Apart from these particular cases, the flow of viscoelastic liquids in
long tubes is generally three-dimensional, i.e. all three components of
velocity, Vi = Vi(XZ, X3) (i = 1, 2, 3) are developed in flow; the occurrence of
transversal components Vz and V3 relating to a secondary flow. Until now
the secondary flows were calculated only for slow tube flows of elastic
liquids 6 ,93 using the evident asymptotic procedure. The pattern of such a
flow in an elliptical tube (as shown in Fig. 8.2) is calculated by using a

Flows of Viscoelastic Liquids in Channels and Pipes

227

Walls

(b)
Fig. 8.2.

(a)
Secondary flows in (a) elliptical and (b) square tubes induced by axial
pressure drop.4

fourth-order approximation 308 (see also Refs 309 and 310). For flows
with higher values of De, these approximations are invalid and analytical
methods of calculations are unknown.
Some experimental evidence for polymer solutions311-313 confirmed
the existence of weak secondary flows with long longitudinal periodicity,
but there was no strict experiment for a high De flow of polymer melts
in a tube with the cross-sections different from plane channel and circular
ones. Some indirect observations did not display any major contribution
of a secondary flow (Ref. 7, p. 272; Ref. 202, p. 136). The necessity for strict
experiments in this field is dictated not only by applications for processing needs, but also the fact that they could also throw light on the validity
of viscoelastic constitutive equations.

8.3 PREDICTIONS OF MAXWELL-LIKE CONSTITUTIVE


EQUATIONS FOR STEADY FLOWS IN LONG TUBES OF
ARBITRARY CROSS-SECTIONS
The above general approach is now applied to viscoelastic constitutive
equations of a differential type. We focus here on the general single-mode
constitutive eqns (2.13) written in the form:
8t+ I/! (c) =0,

I/!(c)=YO()+YIC+Y2C2,

(1=

-P()+/31C+/32C2

(8.4)

where Yk and 13k are functions of temperature and basic invariants I j of


the configuration tensor c which are assumed to satisfy the conditions of
limit transition to the rest state and be Hadamard and dissipative stable.
We now consider the qualitative predictions of flows in long pipes for
three possible classes of constitutive eqns (8.4).

228

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

(12 =0: Existence of Pure Rectilinear Flow


Straightforward calculations show that the second normal stress differin any shear flow, including the flows in long tubes of
ence (12
arbitrary cross-section, if (and only if) the dissipative term in evolution
eqn (8.4) is proportional to the stress tensor, i.e.

8.3.1

(8.5)

This always holds in the particular case when Y2 = and f32 = 0, as in the
examples of Phan-Thien/Tanner upper-convected models (3.42(a)), new
viscoelastic model (5.85) with neo-Hookean potential (a = 0), and also for
FENE (eqns (3.40) and (3.41)), and Larson (eqns (3.43) and (3.44)) models
where the elastic potential depends only on 1 1
We now show that in this case, the problem related to the solution of
eqn (8.2) with a non-slip condition, has an elegant variational formulation
for conditional extremum:

l1>{u} =

Inr G(y)dw=min,
u

where
G(y) =

ulan=O,

f: (J12(~)d~,

y=IVul

udw=Q

(8.6)

(8.7)

o is the cross-sectional domain bounded by the boundary ao;


W= {X2' X3}; dw=dx2dx3; Q is the flow rate, and (J12(y) is the flow
curve in simple steady shearing. If the new functional is introduced as
follows:
11>* = 11> + qQ

(8.8)

where q is a Lagrange's multiplier, the variational problem is reduced to


the Euler equation (8.2); q being the negative pressure gradient. It is
proved that the functional 11> has a positive second variation (or, more
general, is concave) if the flow curve is not decreasing. It should be noted
that this is the stability condition for shearing flow. If the condition is
satisfied, the straightforward method of calculus of variations can be
employed to find the minimizer uo(w). In this case, we can introduce a set
M of the smooth functions v(w) such that
(8.9)

Flows of Viscoelastic Liquids in Channels and Pipes

229

and find the minimizer, uo(w) as the solution of the problem:


'v'v(w)

M:<I>{Qv} ==

Inr G(Q IVvl) dw = min

(8.10)

This solution uo(w) exists for sure if the functional <I>{Qv} is concave. If
the minimizer uo(w) has been found, the pressure gradient q can be easily
determined too. Integrating eqn (8.2) for the minimizer over the crosssection n and using Green's theorem yields:
(8.11)
Here Inl is the cross-sectional area, n is the unit normal vector of the
boundary on with elementary length dv, and t1(y) = 0' dy)/y is the nonNewtonian viscosity. Equation (8.11) establishes one-to-one dependence
q(Q), if the assumptions made about the flow curve are fulfilled.
Thus in this case, the steady rectilinear flows do exist; though, in
contrast with the experimental data, the constitutive equations used
predict the second stress normal difference.
8.3.2 10'21 ~ min (10'121, O't): Weak Secondary Flow Imposed
on a Strong Rectilinear Flow

Consider, for instance, the constitutive eqns (5.85) with the Mooney
elastic potential, with the positive parameter 0( being very small. For this
case, we can consider an approximate asymptotic solution of the steady
flow problem in long tubes represented in the formal power series of
parameter 0(:

0'1'

V2=--,
OX3

P=PO+O(Pl +0(2p2 + ... ,

(8.12)

Here 'I' is the transversal stream function and all the functions in eqns
(8.12), except Po, depend on the cross-sectional coordinates X2, X3' The
'zero' approximation can be found to satisfy eqns (8.2) and (8.3) as the
solution of variational problem (8.6), (8.7) and boundary non-slip conditions. Then the first approximation can be found relatively easily from a
linear set of equations.

230

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Thus in this case, the flow in long tubes of arbitrary cross-section is


almost rectilinear, with a weak secondary flow.
8.3.3 Strong Secondary Flows
If viscoelastic constitutive equations do not predict very small second

normal stress differences, a strong secondary flow (with velocities comparable with those in the basic flow) in long tubes of arbitrary crosssection generally occur at high De numbers. This is, for example, a
common situation for the Giesekus (eqns (3.45)) and specific Leonov
(eqns (1.9), (1.47) and (1.55)) viscoelastic models.
The above classification of the viscoelastic constitutive equations
relative to the secondary flows gives the opportunity to make a proper
choice between these equations using a qualitative conclusion about the
strength of secondary flows observed in experiments.
Assuming that the case described in Section 8.3.2 is the most realistic, it
is possible to introduce a procedure to improve the predictive ability of
constitutive equations. As an example, demonstrated below is a possible
way of improving predictions of the specific Leonov model described by
eqns (1.9), (1.47) and (1.55). In this model, we have introduced the
dissipative term l/I(e) in the evolution eqn (8.4) by reasons of flow-induced
anisotropy in viscosity, the existence of pure planar flows and simplicity.
It seems that the dependence a(e), which was taken ad hoc in neoHookean form, needs to be improved here. In so doing we will retain the
dissipative term as it is, and to improve the dependence a(e) will use the
following two-step procedure.
(a) 1st step
Finding a general dependence aO(e) allowing for
the dissipative term l/I(e) is taken in the form:

a~=O

in simple shear if

2yo= -1,
(8.13)

Using now the Finger formula (1.24) for an elastic potential WO and the
conditions (8.5) results in the relations:
2awo /131 1= R(I 2 + 211 )/6,

(8.14)

The compatibility condition for the elastic potential WO results in the


equation for the unknown function R(Il, 12 )
(8.15)

Flows of Viscoelastic Liquids in Channels and Pipes

231

whose general solution is of the form:


(S.16)

R = f(I) exp( - I t/3),

where f(I) is an arbitrary (supposedly analytical) function of I. Now, due


to eqns (S.14) and (8.16), it is easy to find the general expressions for the
elastic potential W
Wo=
and its two derivatives

2W 1 =

f(z)dz

211+12
6
f(I) exp( - I d3),

(a= 15/e)

2W~= -!f(1) exp(-ld3)

(S.17)

(S.lS)

where I is defined in eqn (S.16). Along with the Finger formula for (J"(e)
and the formulation of the dissipative term "'(e) in the kinetic eqn (S.13),
eqn (S.lS) defines the stress tensor, which satisfies the condition 0"2=0 in
simple shear. This guarantees the existence the rectilinear steady flow in
long tubes of arbitrary cross-section.
(b) 2nd step
Now we define the total elastic potential W as follows:
W(11,1 2) = (1-a)W(11>I z )+a(12- 3)

(8.19)

where WO is the elastic potential defined above and a is a small positive


constant. The smallness of parameter a guarantees that the secondary
flow in the tubes of arbitrary cross-section will be weak enough.
(e) Example
Consider, for instance, in eqn (S.17), fez) =
(S.19) results in:

AZh.

Then, using eqns (S.17)-

/L(l-a) {
(211 +lz +6)n+1
}
W=4(n+1) 1515
exp[-(n+1)(11-3)/3]

+ {la(12 -

(8.20)

3)

where the constant A was defined by transition to the limit infinitesimal


case. The stress tensor is defined as follows:
(J"

+ po = /L(1-a) (

211 +lz+6)n
15
exp[ -en + 1)(11 - 3)/3]

(S.21)

232

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

This formula leads to one-dimensional stable constitutive equations only


when n < - 1. When n = - 1 its predictions are very similar to those for
the neo-Hookean case, but the second normal stress difference will be
predicted very well.

8.4 POISEUILLE FLOWS OF VISCOELASTIC LIQUIDS


IN CAPILLARIES AND DIE SLITS
8.4.1 General Formulation
In this section, we consider the flows in a fully developed region where
the velocity profile is independent from the axial coordinate. To describe
these flows we shall use the multimodal simplified viscoelastic constitutive eqns (5.13) and (5.14) written in the dimensionless form by using eqn
(5.12). These constitutive equations where Yk=O, should be complemented
by the momentum balance equations written in the dimensionless form as
follows:

ov

* 1 0
R-=- -

or

xn

ox

(X n U,'2)

op*
oz*

+P(r)'

P(r)= - -

'

(8.22)
(8.23)

where

v*=v(Jdb,

x*=x/b,

z*=z/b,

n = stands for flow in the die slit of thickness b, and n = 1 for flow in a
capillary with radius b; z and x are the longitudinal and transversal
dimensional coordinates; r:x is the transversal normal component of
dimensionless extra stress tensor, and a! is the dimensionless second
normal stress difference. From now on the asterisk superscript will be
omitted in this section.
Equation (8.23) is integrated to yield:
p= po(z) + p(x),

p(x)=rxx(x)+n

s:

az(xddxdxi'+C(r) (8.25)

where C is a function of time and all variables in eqns (8.22) and (8.25)
depend generally on time t.

Flows of Viscoelastic Liquids in Channels and Pipes

233

8.4.2 Steady Poiseuille Flows


In steady flow, we can roughly estimate the constant C in eqn (8.25) if the
pressure Po in the exit region of the tube and the pulling force 0"0 per
capillary cross-section, acting on the free extrudate, are known. The basic
assumption made in Ref. 202 is that the system of stresses in the exit region
is almost the same as in a region of one-dimensional developed flow, despite
a quick restructuring of the flow field in the exit region. A possible reason
for this may be the small length of the exit zone where the long relaxation
processes which are the main contributors to the stresses, cannot change the
stresses markedly. Under this assumption, the distribution of stresses in a
circular die, following Ref. 202 (Section 5.4.2), is given as:
'ZZ
't(x)= [ 'XZ

P=Po-PZ+'xx+

+ 2fol

o ],

''1''1'

P= Po+O"o
L

'XZ= -xP/2,

C=

'XZ
'xx

J:

(8.26)

0"2 dx/x+C,

[XO"l -0"2(l-x 2 )/(2x)] dx

and in slit die as:


0"= -po+'t(x),

'ZZ
't(x) = [ 'XZ

'XZ= -xP/2,

'XZ
'xx

o
o ],

'yy

P= Po +0"0
L

P=Po-PZ+'xx+ C,

C= Ll O"l(x)dx

(8.27)

We now can use the formulae (5.27) for the dimensionless dependences
of stresses and elastic potential on dimensionless shear rate r. The
dependence r(x) is established by the relation:
- 'XZ = Px/2 = sr +

2rvkllk
Lk l+xk

(Xk =

J ,.---""7
1 + 4(ll kr)2)

(8.28)

234

Nonlinear Viscoelastic Effects in Flows oj Polymer Melts

where Vk and Ak are the rheological parameters defined in eqns (5.12). In


the case s = 0, also useful is formula for the elastic potential W
(8.29)

The velocity distribution 6 and dimensionless flow rate U in capillary


flow, represented as functions of rare:

(8.30)

where t(r) is shown in the right-hand side of eqn (8.28), t w =P/2 and rw
are the values of shear stress and shear rate at the wall, x = 1.
8.4.3 Hardening Phenomena in Steady Capillary Flows 302

Possible hardening phenomena in capillary flows are illustrated using the


single relaxation mode and taking into account the mechanism of fluidity
loss modelled in the previous chapter by shear formulae (7.12). Then the
velocity distribution is of the form:
v(x,P)=

2exp(2/1)

J.b exp(-u)du/u
a

(8.31)

b(P)=4/1/J4-P 2
where the no-slip boundary condition has been used. The dependence of
flow rate U on pressure drop P is given by

and shows that the region of existence of the steady-state flow is restricted
to 0~P~2. If /1=0, P-+2 when U-+oo. If /1>0, the dependence U(P)
defined by eqn (8.32) goes through the maximum; U max and P max
decreasing with an increase in /1. The case P > 2 cannot be considered in
steady flow for the same reason as mentioned in Section 7.5. In the case

235

Flows of Viscoelastic Liquids in Channels and Pipes

f3 > 0, before the maximum U max is reached, the concavity of the velocity
profile is violated. Using eqn (7.12) where now Px/2 stands for x, we
obtain:

~ exp(2f3)
1- ~ 2

exp

(-2f3/ ~).
V ~- " ,

~=Px/2

(8.33)

The condition of violation of concavity is: dr/dx = 0, which is the local


quasi-stationary stability condition for simple shear. It occurs first at the
wall; then x = 1 and ~ = P/2. Easy calculation due to eqn (8.33) gives the
equation for the value of P * (depending on f3) at which the violation of
concavity of the velocity profile first starts to appear:
(8.34)
The dimensionless plot of flow rate versus pressure drop for f3 = 0.5 and
0.9 are shown in Fig. 8.3 (where the values of P * are marked by asterisks).
The dependence P *(f3) represented in Fig. 8.4 corresponds to the onset of
one-dimensional infinitesimal instability in the capillary flow of the
viscoelastic liquid.
Thus, in the case of capillary flow, the relaxation transition of the
viscoelastic liquid to the rubber-like solids proposed in Section 7.5 results
in the instability due to the violation of concavity in the velocity profile.
This fact has been discussed in the literature devoted to the melt fracture
phenomena. 314
0.200

0.150

U 0.100

...-,.

0.050

0.000
0.000

0.500

1.000 P

1.500

2.000

2.500

Fig. 8.3. Dependence of flow rate on pressure drop for flow in a circular pipe:
curve a, /3=0.9, one mode; curve b, /3=O.S, one mode. Onset of violation of
concavity is marked by *.302

236

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

1.900

1.700
Per
1.500

1.300
0.000

L-....~_~-'-~_

0.250

_-'-~~~----'-_~~--.J

0.500

0.750

1.000

{J

Fig. 8.4.

Plot of critical pressure drop (as determined by loss of concavity)


versus f3 for flow in a circular pipe. One mode calculations.302

8.4.4 UNSTEADY CHANNEL AND CAPILLARY FLOWS


OF VISCOELASTIC LIQUIDS
A number of references have been published (see, e.g. Refs 315-323) on
the theoretical and/or numerical studies of the start-up flows of viscoelastic liquids in capillaries and channels under either given constant
flow rate or constant pressure drop. Mostly these works investigated the
Oldroyd, Maxwell, second-order and linear viscoelastic models. Reference 324 also numerically studied the channel flows of viscoelastic
liquids, employing eqns (5.13) and (5.14) with 1'k=0. In this paper, some
comparisons were also made between the results of numerical calculations and data, including the measurements of stress by the birefringence method.
Some results of numerical simulations of hardening phenomena in
unsteady Poiseuille flows of viscoelastic liquids were obtained in Ref. 302
using the multimodal extension of eqns (7.11) and (7.12).
Two relaxation modes with an additional viscous term were used in
calculations in Ref. 324. The parameters of the model
'11> '12=3.58 x 104 , 2.95 X 104 Pa s-\
()b()2=6.07,0.47s;
s=0.01

(8.35)

were evaluated by fitting the steady dependences of the flow curve and
the first normal stress difference on the shear rate for polyisobutylene
Vistanex LM-MH (Enjay) at 27e.

Flows of Viscoelastic Liquids in Channels and Pipes

237

Two basic regimes of channel flow were analysed numerically in Ref.


324.
(i) Flow rate U(t) was given as a step-wise function smoothed over a
very short period of time. In this case, the inertialess approach
is valid and the problem was reduced to the inertialess inhomogeneous simple shearing with a linear gapwise distribution of the
shear stress. The results of calculations for gapwise distributions of
velocity, shear rate, shear stress, and for the first and second normal
stress differences are shown in Figs 8.5 and 8.6. The calculations of
the time evolution of pressure drop shown in Fig. 8.7 displayed the
occurrence of a maximum whose intensity increased with the
increase in U. In Ref. 302 where the calculations of this regime were
also performed including the inertia term in the momentum eqn
(8.22), small-amplitude decayed oscillations were observed.

(b)

(0)

0.3

,....

'
..

0.2

15

....... 10

11

.)0..

><

:::J

0.1

Fig. 8.5. (a) Predicted gapwise distributions of velocity u at various times t (in s)
after imposing a constant bulk velocity U = 1.72 x 1O- 3 m S-l: t=(1) 0, (2) 0.02, (3)
1.0 and (4) 10. (b) Predicted shear rate at various times t (in s) after imposing a
constant bulk velocity U=1.72x 1O- 3 ms- 1 : t=(1) 0, (2) 0.02, (3) 0.1, (4) 0.4, (5)
1.0, (6) 3.4, (7) 5.0, (8) 10.0; b=2 x 1O- 3m (8=2 x 1O- 4 m).324

238

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

u = 1.72
......

x10- 3 m 5-1

2.6

10

'"'E
t(s)

7
U = 1.72 x 10- 3 ms-1

4.0
10

a
......

'"'E
Z

v,

"
'"
1---

1.0

1.0

'",2

"
z
0.4

0.4

0.1

3
4

U= 1.72 x10- 3 m
0.2

......

'E

0.1

t (s)

5-1

4.0
10

1.0

'-'"

<t

'0

0 ..02

00

0.5

y/b

"

'"

1.0
0.4

0.5

y/b

0.2
1.0

Fig. 8.6. (a) Predicted gapwise shear stress '12; (b) first and (c) second normal
stress differences N 1 and N2 at various times after imposing a constant bulk
velocity U; b=2x 1O-3 m (e=2x 1O-4 m).324
(ii) Pressure drop P(t) was given as a step-wise function smoothed
over a very short period of time, as proposed in Ref. 325. In this
case, the inertialess approximate solution for flow rate U(t) hardly
differed from the exact one after 1O- 4 s. Figure 8.8 demonstrates
the gapwise distributions of the velocity and shear rate, whereas
Fig. 8.9 represents the distributions of first and second normal
stress differences obtained in these calculations.
The results of the calculations in this regime, shown by solid
lines in Figs 8.10 and 8.11, were compared in Ref. 324 with
experimental data shown by symbols. Figure 8.10 displays the

Flows of Viscoelastic Liquids in Channels and Pipes

239

tCS)

10

Fig. 8.7. Dependence of predicted pressure gradient P on time after imposing a


constant bulk velocity U for e=() lx1O- 1 m, () 2xl0- 4 m, () 5x1O- 4 m;
b=2 x 1O- 3 m. 324
(0)

1.5
4

20
"i

oil

1.0

'"'

..

'"Q

.....
"i

:J

oil

0.5

10

.>8

1.0
0~------~~-------3

Fig. 8.8. Predicted gapwise distribution of (a) velocity u and (b) shear rate y at
different times after imposing a pressure gradient trace with a final value of
Po = 3.86 X 10 7 Pa m -1 as given in Fig. 8.7. t(s): (1) 0.005, (2) 0.01, (3) 0.02, (4) 0.05,
(5) 0.1, (6) 0.2, (7) 0.5, (8) 1.0, (9) 2.5, (10) 5.0, (11) 10.0.324

240

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

t(s)
10

t(s)
10

,..
0

C1.

.,.,

,..
0

C1.

'0

'0

"
Z

"N

2
0.5

0.5

0.2
0.1

0.2
0.1
0

0.5

y/b

1.0

Fig. 8.9. Predicted gapwise distribution of (a) first and (b) second normal stress
differences N 1 and N 2 on time after imposing a pressure gradient trace given by
Ref. 325; b=2 x 31O- 3 m. 324

gapwise distribution of birefringence, dn = CJ crr + ai2; the value


ofthe stress-optical coefficient C = 1.414 x 10 - 9 being measured in
Ref. 326. Figure 8.11 demonstrates a comparison between theoretical predictions and experimental results for the time dependence
of the flow rate U(t}. This dependence was experimentally obtained by measuring the displacement of the plunger. Though the
predicted and measured dependences are in qualitative agreement,
they sometimes differ by more than an order of magnitude despite
the good agreement between the results for birefringence. No
analysis of such heavy discrepancies was given in Ref. 324.
Perhaps they are related to such factors as inertia of the plunger
and the compressibility of the polymer in pre-capillary flow, which
were not taken into account in these measurements. Additionally,
the large-amplitude decayed oscillations on the dependence U(t}
were found in calculations. 302 Surprisingly enough, in this highly
nonlinear region of deformation, the frequency of the oscillations
coincided well with that predicted by the linear theory of viscoelasticity.

241

Flows of Viscoelastic Liquids in Channels and Pipes

1.0
, =2 5

'=0.5 s

, = 105

0.5

...Q

0
1.0

, = 1.0 5

<l

, =5 5

0.5

0.5

1.0

y/b

00

0.5

1.00

0.5

y/b

y/b

1.0

(a)

...

c::

<l

:~

..

t =6.5 5

t= 1.0 5

0
0

<l

00

1.0

1.0

(b)

Fig. 8.10. Predicted (solid lines) and experimental (points) gapwise birefringence
~n distribution at different times after imposing a smoothed pressure gradient
with a final value of Po=(a) 1.27 x 10 7 Pam- 1 and (b) 2.82 x 10 7 Pam-l;
b=2 x 1O- 3 m. 324

242

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

-I

-6--6 2.B2 xlO 7

Fig. 8.11. Predicted (solid lines) and experimental (points) dependences of bulk
velocity U on time after imposing a smoothed pressure gradient with different
final value Po. 324

Two results of simulations of the hardening phenomena in Poiseuille


flows of viscoelastic liquids 302 are demonstrated for the single-mode
approach in Figs 8.12 and 8.13. Figure 8.12 shows the increase in the
pressure drop when a given constant flow rate U exceeds the critical
value. This behaviour is similar to that found in Section 7.5. The
slow-down effect in the Poiseuille flow of viscoelastic liquid is demonstrated in Fig. 8.13. No experiments, to the authors' knowledge, were
undertaken to verify these predictions.
8.4.5 Enhancement of the Flow Rate Using a Pressure Drop
with a Pulsatile Component
Consider the isothermal extrusion of polymers through a slit or circular die by imposing a pulsatile component of the pressure at the die
entrance:
(8.36)

where the amplitude factors pO and B (dimensionless) and frequency (J) are
constant. The mean values of the pressure <p) and flow rate <Q)

243

Flows of Viscoelastic Liquids in Channels and Pipes

(3=0.5.

0=0.152

10.000

7.500

5.000
P
2.500

0.000

10

15

20

Fig. 8.12. Dependence of pressure drop on time in a circular pipe when the flow
rate is given. One mode, supercritical case with inertia term taken into account. 302
0.400 r - - - - - - _P_=_I_.9_9---,{3_=_0_._1_ _ _ _ _ _ _~

0.300

IT
,0.200

Fig. 8.13. Dependence of flow rate on time in a circular pipe when the
pressure drop is given. One mode, supercritical case. The curve has been
smoothed, so it does not show the oscillations due to the inclusion of the inertia
term.302

averaged over the period of oscillations T, T = 2n/OJ, are defined as


follows:
1
<p)=;y

fT Po(t)dt=po,
0

<Q)=-

fT Q(t)dt

(8.37)

We can also define the amplifying flow rate factor S(6, OJ) as
S=<Q)/Qo-l

(8.38)

244

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where the flow rate Qo corresponds to the constant component of


pressure pO.
Experimental and theoretical studies327-335 revealed an especially high
positive effect (s> 0) in extrusion, when the pulsatile component of the
entrant pressure is large in amplitude and low in frequency. Though for
high pressures this effect also employs the density variations,330-332 we
consider below only the major, incompressible case, demonstrating that
this effect is caused by the shear thinning. This means that it can be
applied not only to pure polymers and their solutions but also to
dispersed systems, compounds, etc.
Consider the simplest one-dimensional case of quasi-steady flow in the
die slit of thickness h and width b. Then, neglecting the entrance losses,
we can write the familiar expressions for pressure, p(x, t), and the shear
stress, ,( y, t), distributions as follows:
p=p0(1 +e sin wt)(L-x)/L

(8.39)

,= -p0(1 +e sin wt)(h/2L)(1-2y/h)

(8.40)

where the x-axis is directed along the centreline (y=h/2) of the gap. For
low-frequency oscillations (Ow ~ 1) non-steady effects are negligible and
only the steady-state flow characteristics of the polymers should be used.
In most practical cases, the 'power law' approximation of the flow curve,
y= com, is valid over a wide range of shear rates. Then, using the common
expression for flow rate - pressure drop dependence,
Q=a

laxapl

'

(8.41)

and eqns (8.37)-{8.39), we can represent the expression for the amplifying
factor s as follows:

W f21t/(f)

s(e, w, m) = 21t

(1 + e sin wt)m dt - 1

(8.42)

For example, s=O, e2 /2 and 3e 2 /2, when m= 1, 2 and 3. Equation (8.42)


shows that amplifying factor s is positive for shear thinning (m> 1) and
negative for shear thickening (m< 1).
This effect can also be qualitatively understood from the inelastic
(instant) response for shear rate y on the harmonic oscillations of shear
stress, (Fig. 8.14) with a constant component '0 (1'01 >0). Due to the
highly nonlinear character of the flow curve shown in Fig. 8.14, the cycle
of y oscillations is asymmetric, with the average value of the shear rate

Flows of Viscoelastic Liquids in Channels and Pipes

1.00

Wi

0.50

245

01 I 05 .
0.25~ 11:0 i
0.50

~i+-=----+-I

0.75~ ~
looL
.

1:1

wt

Fig. 8.14. Nonlinear distortions of oscillations in inelastic nonlinear media: 1,


flow curve; 2, curve of function T=To(1+e sinwt); 3, curve of function y(t)
corresponding to curve 2.

being higher than that corresponding to the average stress '0. This effect
is, in principle, similar to the nonlinear amplification of electric oscillations in the vacuum-tube amplifier having nonlinear transfer characteristics.
As seen from eqn (8.42), the amplifying effect increases with an increase
in amplitude oscillations e and power index 'm' of the flow rate. That is
why it is wise to apply this effect to polymers with essentially nonlinear
flow curves, such as PVC, PP, PS, PE, thermoelastoplastics, rubber
compounds, etc. 333 To make the 'vibroforming' effect more pronounced,
the parameters of the vibrations should be chosen so that the amplitude
of the periodic shear rate would not be less than its time-averaged value.
The practical and design aspects of vibroforming are discussed in detail in
Ref. 333.
At arbitrary frequencies, where the shear elasticity of polymers is
important, the extrusion was experimentally and theoretically studied in
Refs 327, 328, 334 and 335. The results of Refs 327 and 328 showed that
the amplifying factor s has a maximum at a certain frequency w. It was

246

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

also shown 334 that at a fixed frequency, the amplifying factor s reaches a
maximum at a certain constant pressure pO.

8.5 ENTRANCE AND EXIT STEADY FLOWS


8.5.1 Experimental Results
From the early 1970s onwards, a lot of studies were concerned with the
flow of polymers in so-called 'conjuncture' regions, i.e. in the regions with
abrupt changes in the geometry of channels.336-353 The classical example
of this problem is the flow in a pre-capillary region, named a contraction
flow, when the cross-section abruptly contracts downstream. The contrary example is the flow in an expansion region when the tube crosssection abruptly expands downstream. Besides the capillary rheometry,
these problems have a lot of applications in polymer processing, especially in extrusion and injection moulding. The experimental and numerical
results reported in the literature mostly discuss the contraction axisymmetrical steady flows of polymers in the plane and circular tubes. Various
experimental techniques were used in these studies: stroboscopic visualization and laser-doppler anemometry for detecting the streamlines and
velocity profiles, and birefringence to find the stress distributions in the
flow.
Almost all polymer solutions display vortices in contraction symmetrical steady flows in plane or circular pipes except for very slow flows. As
compared with polymer melts, the vortices are more pronounced for the
flow of polymer solutions. Here the Reynolds number, Re, may also play
an important role. Figures 8.15-8.17 taken from Ref. 352, demonstrate
two such flows of a 'Boger' fluid: dilute solutions (~0.1 %) of polyacrylamide in a maltose syrup/water solvent, for a very slow flow when De -+ 0
(Fig. 8.15), and flow with a finite value of De (Fig. 8.16). Figure 8.17
clearly demonstrates the effect of tapering on one side of the entrant
channel. It has also been reported 353 that in planar contraction flows of
polymer solutions with intermediate values of De and a high contraction
ratio, two vortices occurred sometime at each side of the half-symmetry: a
smaller 'lip' vortex near the entrance, and a larger 'corner' vortex at the
upstream tube corner.
However, from the very beginning some confusing results were obtained for the contraction flows of polymer melts. It was found that for
some polymers, vortices in pre-entry flows occur, starting from a certain value of De and growing as De increases, but for other polymers,

Flows oj Viscoelastic Liquids in Channels and Pipes

Fig. 8.15.

247

Streamlines for slow flow of Boger fluid B62 in a 16:1 square/square


contraction; We= 1.8 x 10- 6 352

they do not occur (at least for the same values of De). For example, the
vortices were detected for LDPE,336-350 PS,336.347 PMMA,336.338
branched PDMS,340 and pp.336 However, they were never found for
HDPE,13.336.337.342.348.351 PB,346.347 PI,347 and linear PDMS. 340
The answer to this fundamental question was first proposed by
Cogswell. 338 Based on the assumption of the dominant role of elongation
flow in the entrance region, he found that the formation of the vortices in
the contraction flows of polymeric melts is tightly related to the behaviour of the viscosity '1e1 in simple elongation: the vortices exist only when
'1e1 increases with a growing elongation rate and do not occur otherwise.
This correlation was further confirmed in Ref. 348 where the formation of

248

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Fig. 8.16.

Streamlines for fast flow of Boger fluid B60 in a 16:1 square/square


contraction; We "= 9 x 10- 4 ,352

Fig. 8.17.

Streamlines for fast flow of Boger fluid B60 in a 16:1 square/square


contraction with one tapered lip; We=9 x 10- 4 ,352

Flows of Viscoelastic Liquids in Channels and Pipes

249

the vortices was considered as a stress relief for extensional entrance flow,
and recently in Ref. 350.
It should be mentioned that from a rheological viewpoint, this result is
fundamental, because it demonstrates that the nonlinear rheological
behaviour of viscoelastic polymer systems is not predetermined only by
their linear viscoelastic properties, but also depends on some additional
parameters which are displayed only in the nonlinear region of deformation. Some examples of such behaviour are demonstrated in Chapters 6
and 7.
It is known, and stressed once again in Ref. 350, that sometimes the
difference in elongational rheology can be achieved for the same polymer
by a variation in the MWD. For instance, Miinstedt 354 found that
unimodally distributed PS shows the bounded stress in simple elongation, but that bimodally distributed PS shows a sharp increase in
elongation viscosity. Thus one can believe that the entrance flows of two
PSs will be completely different. Unfortunately, advances in molecular
rheology have not yet been enough to predict this type of behaviour, and
for a realistic modelling of viscoelastic constitutive equations, one needs
to anticipate the modelling of this effect at least on the phenomenological
level. The predominant role of elongational behaviour on the contraction
flow can be also seen in the fact that some polymers (e.g. PS)350 display
the occurrence of vortices in axisymmetric flow and do not in the planar
flow.
Additionally, unlike the low Re flows of inelastic liquids, the contraction and expansion flows of viscoelastic liquids are different in both
kinematics and stress distributions. 13 ,355
In order to estimate, at least qualitatively, the difference between
Newtonian and viscoelastic (Re -+ 0) contraction flows in the entrance
region, let us briefly discuss some patterns of their velocity distributions.
Though these profiles can be obtained using tracers, they are measured
most precisely by using laser-doppler anemometry. Paskhin 181 was
seemingly the first to apply the laser-doppler technique for measuring
complex flows of viscoelastic fluids. Some results were also obtained in
Ref. 356.
Figures 8.18 and 8.19 181 demonstrate the longitudinal and crosssectional distributions of the longitudinal velocity u(x, y) in the entrance
region of a contraction flow for Newtonian (Fig. 8.18) and viscoelastic
(Fig. 8.19) fluids. In these experiments, channels of square cross-sections
were used. Divinyl-nitrile rubber served as a Newtonian liquid and 40%
butyl rubber solution (M w ~ 3 x 104) in transformer oil, with a maximum

250

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

-2

Ca)

(b)

Fig. 8.18. Variations of dimensionless longitudinal velocity U o(X, Y) for Newtonian fluid in contraction flow at different mean velocities Uo m2 s -1 in smaller
channel. (a) Variation of centreline velocity U o(X, 0) along the channel axis
X: (0) Uo =0.63; (6) Uo = 1.73; (x) Uo = 5.2, (0) Uo = 10.1. (b) Cross-sectional
distributions of longitudinal velocity for various cross-sectional positions X,
uo=1.73: (0) X= -0.29, (0) X= -0.1, (6) X=OA, (0) X=3.181

Newtonian viscosity '70 = 1.7 X 104 poise, was used as the elastic liquid.
Unlike the Newtonian case (where the centreline velocity increases
monotonically in the entrance region), in the viscoelastic liquid, a
maximum in the centreline velocity starts to occur at higher De, which is
intensified and moves towards the entrance line with the increase in De
value. Also, Fig. 8.20 manifests the difference between the Newtonian and
viscoelastic flows in the exit region of the channeJ.181 It is noticeable that
the centreline velocity in the Newtonian flow monotonically decreases
whereas for the viscoelastic flow, it goes through a maximum.
8.5.2 Numerical Simulations

There is a large list of publications devoted to numerical simulations of


contraction flows of viscoelastic liquids. In the calculations therein,
two-dimensional planar or axisymmetric steady problems were usually
considered for the closed set of equations consisting of the constitutive
equations, the momentum balance equations (usually taken within the
inertialess approach), and the incompressibility condition. The great
majority of the calculations were focused on the constitutive equations of
differential type, which are more flexible and easy to use in calculations.

251

Flows of Viscoelastic Liquids in Channels and Pipes

0.9.-----~------~----~

20

-1

40

(a)

(b)

(e)

Fig. 8.19. Variations of dimensionless longitudinal velocity U o(X, Y) for viscoelastic fluid in contraction flow at different mean velocities uo(m 2 S-1) in a
smaller channel. Variation of centreline velocity U o(X, 0) along the channel axis
X: (a) in the inlet region of the channel and (b) in the pre-inlet region of the
channel: (0) uo=0.67, (.6.) uo=2.92, (<) uo=7.77, (0) uo=20.1, (T) uo=46.6. (c)
Cross-sectional distributions of longitudinal velocity for various cross-sectional
positions X, uo=2.28:(<) X= -0.6, (0) X=0.3, (.6.) X=1.4, (0) X=20. 181

Two major problems arise here: the proper choice of constitutive


equations and the proper choice of a numerical scheme. The first problem
is associated with the analyses of stability of the set of equations which
have been discussed to some extent in Chapter 3. As stated in this
chapter, the constitutive equation chosen for calculations should be
Hadamard and dissipative stable and should have monotonically and
unbounded increasing flow curves in both simple shear and simple
elongation. If these conditions are not satisfied, heavy instability could

252

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

u
0.8

0.6

-2 X-L
(a)

0.9

-L...-----'---~0.8

-4

X-L

-2

(b)

-1

(e)

Fig. 8.20. Variations of dimensionless longitudinal velocity U o(X, Y) for


Newtonian and viscoelastic fluids in exit flow at different mean velocities
uo(m 2 S-l). (a) Variations of centreline velocity Uo(X, 0) for Newtonian fluid
along the channel axis X: (0) uo=0.54, (L.,.) Uo= 1.65. (b) Variations of centreline velocity Uo(X, 0) for viscoelastic fluid along the channel axis X: (0)
uo=0.296, (L.,.) uo=0.96, (0) Uo = 2.68, (D) uo=7.35, (v) uo=20.2. (c) Crosssectional distributions of longitudinal velocity for various cross-sectional
positions X1=L-X, uo=2.68: (0) X 1=12, (L.,.) X 1=1.2, (D) Xl =0.3,
(0) Xl =0.06. 181

occur even at very low values of the De number. However, even if these
conditions are met, they do not guarantee us from the onset of a two- or
three-dimensional fluid mechanical instability of flow, as observed in
experimental studies.
On the other hand, the numerical procedure chosen for the calculations
can produce numerical instabilities of different types. Any numerical
method starts with the procedure of discretization of nonlinear partial
differential equations which form the closed set. Two methods are in use
in numerical analyses of viscoelastic flows: the finite difference method
(FDM) and the finite element method (FEM).357 FEM is the most
popular nowadays because it is more flexible in the discretization of
equations near the corners and borderlines. The procedure commonly
used in FEM is to subdivide the total domain into subdomains or
'meshes' of different sizes and configurations convenient for analysis, and
representing in every mesh the space distribution of all the functions
searched for by easy polynomials satisfying some smoothness conditions
on the mesh boundaries. Then, using the Galerkin procedure, the set of

Flows of Viscoelastic Liquids in Channels and Pipes

253

nonlinear partial differential equations is reduced to the nodal set of


either ordinary differential evolution equations for an initial boundary
problem, or to the set of nonlinear algebraic equations for boundary
problems of steady flow. This set is then solved using a proper 'fixed
point' method, e.g. the Newton-Raphson iterative scheme. If for a used
mesh configuration, this method converges for given conditions of flow
(e.g. for a given value of De), then one should be convinced that the
solution is mesh independent, i.e. convergent for a finer set of meshes.
There are few mathematical results to proove the convergence of the
calculations and this is practically achieved by comparing two solutions
after mesh refinement. In applications of the numerical simulations of
viscoelastic flows, these methods and their modifications are discussed in
Refs 358 and 359.
The numerical problems occurring in simulations of high De viscoelastic flows are discussed in detail in Refs 360-362. The common
numerical strategy in obtaining the results is described as follows. The
first step in the calculations is the solution of the Newtonian problem to
obtain an initial estimation for the velocity field. This estimation is then
used in the straightforward calculations of the algebraic set of equations
for incrementally increasing values of De. Usually, the calculations are
continued up to the point of divergence of the numerical scheme with the
highest value of Deborah number, De*, achievable in calculations. The
value of De* obtained in these calculations may not, however, exceed
several units, depending on the choice of constitutive equations, and is
enhanced by introducing a Newtonian term into the constitutive equations. The first numerical instability appears at the lip corner of the
entrance region and displays heavy oscillations near the corner. It has
been shown 360 that the instability increases near De=De* with mesh
refinement, which, in turn, diminishes the value of De*.
Another, more sophisticated, numerical strategy was applied in Ref.
355 to solve the problems of steady planar contraction and expansion
viscoelastic flows in a multimodal Leonov model with an additional
Newtonian term. The complete set of equations is:
-Vp+SI1V2V +

L l1ek v.ck =O
k

(8.44)

V.V=O
t

(8.43)

V. VCk-Ck' VV -(VV) ,Ck+ 2e k (Ck -0)=0


2

(8.45)

254

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Here the multimodal set of constitutive equations (1.71)-(1.76) was


employed with nk = Yk = 0 where Ck(X) are two-dimensional matrices
represented in eqn (5.1), p(x) is the pressure, V(x) is the two-dimensional
velocity vector, x = (x b X2) is the point in the flow domain, and s, 1], 1]k
and (}k are rheological parameters considered as constants in the isothermal flow. Their values are usually estimated in linear rheometric
experiments. Equation (8.43) is obtained after substituting the expression
of shear stress through the tensors Ck and strain rate tensor into the
inertialess momentum balance equation. Equation (8.44) is the continuity
equation, and eqn (8.45) is the evolution equation for the tensor Ck in the
particularly specified Leonov model. Equations (8.44)-(8.45) form the
closed set which are solved under non-slip boundary conditions.
The following three-step numerical scheme was proposed in Ref. 355.
(1) Estimation of the primary velocity field from the numerical solution of the Newtonian problem.
(2) Solution of eqn (8.45) for the given value of De with a known
primary velocity field to estimate primarily the tensors Ck' This is
the key step in the method. The values of Ck in this step are found
as the solution of the Cauchy problem for eqn (8.45) searched for
along the streamlines of the primary velocity field, i.e. by using the
stream-wise procedure with an initial value of c~ specified upstream.
(3) Solution of eqns (8.43) and (8.44) with known tensors Ck found in
the previous step, in order to estimate new values of velocity field
and pressure. A finite-element formulation with four successively
refined sets of triangle meshes in the entrance region was used here.
These iterative procedures were repeated until velocity V and Ck fields are
stabilized with a given accuracy. To stabilize the computations at high
De, the artificial term wV2(vm - vm - 1) is added to eqn (8.43), where vm
and vm-l are the velocities at the current and previous iteration and w is
a constant. Thus eqns (8.43) and (8.44) at the mth iteration take the form:

vvm=o

(8.48)

Obviously, eqn (8.46) coincides with eqn (8.43) at the convergence where
Vm - 1 ~ Vm Numerical experiments have shown that the chosen value of
w should be close to 1 to produce the most stable results.

Flows of Viscoelastic Liquids in Channels and Pipes

255

The numerical scheme demonstrated the steady results in solving eqns


(8.43)-(8.45) up to the values of De::::: 270 (!). This is two orders of
magnitude higher as compared to the results of numerical calculations
reported previously in the literature.358-362 Surprisingly enough, these
extraordinary numerical results have never been discussed, rejected or
confirmed by other scientific groups specializing in numerical computations of viscoelastic flows!
It is also remarkable that the above numerical scheme 355 when applied
to the upper-convected Maxwell model failed to converge at De::::: 3. This
result may be directly related to the dissipative instability of the model in
the elongation flow dominated in the contraction flow near the entrance
region. No results were reported in Ref. 355 for the numerical computations of the set (8.43)---(8.45) in the particular case s = 0, i.e. without the
Newtonian stabilizing term. However, in this case, one can expect a
numerical instability at lower values of De due to the dissipative instabilities described in Chapter 3. These examples clearly indicate the importance of the nature of constitutive equations on the numerical stability.
The problem of numerical computations of viscoelastic entrance
flows without stabilizing Newtonian terms in constitutive equations is

1.5
000 0 0 0 00 0 0 00 00 0

0,,"" " " " " " " " " " " "" " " "
XX)( )(

x)()(

o
o

x)()(

)(

De: a

"De:

x )( )(De=50

u/U

)( ) ( ) ( ) (

1.0

)(

')(

0.5

-8

Fig. 8.21.

-4

x/b

The dimensionless longitudinal velocity u/U along the symmetry line


for De = 0, S and soY

256

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

(a)

Contraction

2.0

Q(cm3 s-1)
00.0272
A 0.0759
[J 0.3080

[J

[J
[J

-5

x/b

(b)

2.0

Expansion
Q (cm3

s-1)

0.0432
A 0.117
[J 0.343
0.955

Fig. 8.22. Birefringence along centreline for (a) contraction and (b) expansion
flows at different flow rates. Lines represent theory and points represent experimental data. 355

257

Flows of Viscoelastic Liquids in Channels and Pipes

seemingly related to the asymptotic analyses of singularities at the lip


corner, which different constitutive equations would demonstrate in
different ways. The same (and even more serious) problems arise in the
numerical computation of viscoelastic exit flows, where the capillary
borderline is the origin of singularities for the velocity gradient and stress
fields. Without analyses of these singularities, even very high-resolution
straightforward numerical methods will diverge at comparatively small
Deborah numbers.
8.5.3 Comparison of Numerical Simulations with Data
The results of numerical computations of entrance flows 13 ,355 allow the prediction of such important characteristics of the flows as the cross-sectional
and axial distributions of the longitudinal velocity and intensity of the
stresses across the gap. In both papers cited, the flow birefringence experiments were undertaken for the PIB Vistanex LM-MH (Enjay) at 27e.
In Ref. 355 the rheological behaviour of the polymer was modelled by
two relaxation modes and an additional Newtonian element with the

1l(-00

(a)

x/b=10.0
Q=0.0272 cm3 s-1
U/b =0.17s-1

t1
2Vl.32 2_
2
y/b

2[

y/b 1

)~~ ~ ~o~~~s-1
o

-8.0

-3.26

-1.36

00

0.32

1.46

x/b=1O.0

;;1 5

Y/b

1l

U/b=0.47s-1 3
Q=0.0759cm

1-1.0-1

(e)

-0.74

Q=0.30Bcm 3 s-1

...

U/b=1.~~S-:

0 .28

~nx104

00

o{

....

~
0

Fig. 8.23. Gapwise birefringence distribution at different cross-sections in contraction flow at various flow rates. Lines represent theory and points represent
experimental data. 3 5 5

258

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

U/b =0.27s-1
Q:0.0437cm3 s-1

.t

x/b=10.0

2~08to~3521(b)~2.0

y/b

0 '

jo ?oQ_=O~'118Cm3S-'
U/b= 0.73 S-1

-00

x/b=10.0

00

2V:l}?
(e)

~b

-u

-2.52

-0.92.

x/b= -0.36,

0 0

0 0

U/b= 2'14 s-1


Q=0.343cm 3 s-1

U/b=5.97s-'
Q=0.955cm3 s-1

(d)

y/b

1- -1
1.0

Fig. 8.24. Gapwise birefringence distribution at different cross-sections in expansion flow at various flow rates. Lines represent theory and points represent
experimental data. 355

values of rheological parameters shown in (8.35). In the experiments,


birefringence data, ~n = C aI + aI2, were measured along the centreline and across the channel, with the value of stress-optical constant
C= 1.414 x 10- 9 . 326 Figure 8.21 shows the predicted dimensionless longitudinal velocity distributions along the symmetry line for the values of De
equal to 0, 0.5 and 50. One can see that the Newtonian case, De = 0,
coincides quantitatively with the laser-doppler data of measurements
shown in Fig. 8.18(a). Also for De>O, the predicted results are in

Flows of Viscoelastic Liquids in Channels and Pipes

259

qualitative agreement with the data for another PIB, shown in Fig. 8.19.
Lines and symbols in Fig. 8.22 demonstrate the results of calculations and
data, respectively, for the birefringence along the centreline at different
flow rates for contraction (Fig. 8.22(a)) and expansion (Fig. 8.22(b)) flows
of PIB. Figures 8.23 and 8.24 show the gap wise birefringence distributions at different cross-sections in contraction and expansion flows,
respectively, and at different values of flow rate. Here the solid lines
represent the computational results, and the symbols show the data. The
difference between the two flows is very noticeable.
More extensive comparisons of the calculations with data for the same
polymer are represented in Ref. 13, where the effect of the contraction
angle was also under study. The results in Ref. 13 show the same
reasonable agreement between the predictions and data as these demonstrated in Figs 8.21-8.24.

CHAPTER 9

N on-isothermal Flows of Polymeric Liquids

9.1 INTRODUCTION
This chapter presents some basic effects of non-isothermality in flows of
viscoelastic polymeric liquids. First, the equation for temperature variations is derived on the basis of the common assumption about entropy
elasticity for polymeric liquids. Though some possible alternative approaches are also briefly discussed, in the following we use the specification of the dissipative heat term resulting from the entropy elasticity
assumption. Then, specifying temperature dependences of rheological
parameters allows us to formulate a closed set of general thermorheological equations valid for the basic viscoelastic constitutive equations employed in this monograph. We also show how the principle of
time-temperature superposition facilitates searching for possible dependences of rheological parameters on temperature.
A number of steady non-isothermal shear flows with substantial
dissipative heat generation are further considered to illustrate the effects
of non-isothermality. These are of particular importance in understanding rheometric studies. For example, the effects of dissipative heating
produce the 'rheological non-invariance', bring additional disturbances in
flow, and sometimes even result in the occurrence of non-steady flows.
Most effects considered for viscoelastic liquids are similar to those known
for Newtonian liquids, which, in turn, are similar to those observed in

Non-isothermal Flows of Polymeric Liquids

261

combustion theory. The main difference between common viscoelastic


shearing flows and those for Newtonian liquids, is that in the former,
normal stresses occur which are usually dynamically insignificant. Yet, in
certain flows, the normal stresses seemed to be extremely important. For
instance, the study of normal stresses in shear flows between rotating
disks, while taking into account dissipative heat generation, throws light
on the operation of sealing for the rotary shaft of a new type, whose
functioning is based on the effect of normal stresses.
There is a remarkable difference between steady shear and extensional
non-isothermal flows, thermostatted on the surface but with substantial
dissipative heat generation. While heat generation does not change the
basic kinematics of shear flow, it changes completely the flow kinematics
in extension where, because of non-isothermality, notable tangential
stresses occur on the surface, even if the specimen is thermos tatted. This
effect is illustrated by experimental data on isothermal retardation after
a uniform extension when the time scales of heating and relaxation are
commensurable.
It should also be mentioned that non-isothermal flows of viscoelastic
liquids are considered in Ref. 363 but for the case when the effects of
dissipative heat generation are negligible. Also, a lot of non-isothermal
problems in polymer processing are considered in Ref. 2.

9.2 EQUATION FOR TEMPERATURE VARIATIONS


This equation was obtained in Refs 364 and 365 for the general case of
simple liquids under the assumption of entropy elasticity in viscoelastic
liquids. It was assumed 23 ,364 that due to the formation of temporary
entanglements in elastic liquids, the change in their conformation entropy s, makes an overwhelming contribution to variations of free
energy f = u - Ts at T= constant, just as in the case of crosslinked
rubbers. The possible contribution of local strains in internal energy u,
related to the distortions of valent angles between monomer units in
polymeric chains, is considered to be negligible in this approach. Therefore this assumption results immediately in the relation: u = u(p, T)
where p is the density. This assumption was also confirmed experimentally for simple shear and simple extension flows of a polyisobutylene
melt. 364 The equation for the temperature variations, obtained in Refs
364 and 365, is also true for the constitutive equations considered in
Chapter 1, because these are a particular case of the general constitutive

262

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

equation for a simple liquid. Now we shall reproduce this derivation for
our particular case. 300
Assuming that the Fourier thermal conductivity relation (1.37) is
fulfilled, we can employ the First Law of Thermodynamics written in
local form as follows (see eqn (A. 54) in Appendix A2):

p du/dt = V KO VT + tr((1e)

(9.1)

where K is (generally) the second-rank thermal conductivity tensor, t is


time and (1 and e are the stress and strain rate tensors, respectively.
According to eqn (A.64) in Appendix A2, the heat capacity C at
constant elastic strains Ck is given by:
CCk == C = T : ; ICk =

:~p

(9.2)

where the entropy elasticity assumption has been used. In the incompressible case, we consider further, u = u(T) and C = C(T). Thus, eqn (9.1)
takes the form:

pCdT/dt= V KO VT +tr((1oe)

(9.3)

Equation (9.3) can be obtained from the Second Law of Thermodynamics (eqn (A.62)) by substituting the expression for the entropy
production (eqn (A.63)) with an allowance for Yk ==Ck.
It is convenient to measure the heat capacity in liquid under relaxation condition, where C(T)=dQ/dT; Q being the amount of heat. The
basic thermo-physical constants, melting temperature Tm, density p,
scalar thermal conductivity K, heat capacity C p (which is equal to C for
incompressible fluid), and activation energy of viscous flow E, are
represented in Table 9.1366-369 for some industrial polymer melts as
compared to those for water. 370 Fairly wide variations for some constants in Table 9.1 are due to the multiplicity of industrial polymer
grades. More precise data on the values p,C p and K, along with their
change with temperature can be found in Refs 366 and 367.
Now let us analyse how the entropy elasticity assumption can affect
the constitutive equations proposed in Chapter 1. According to eqns
(1.17) and (1.18), the total free energy and entropy of the liquid IS
written as follows:

s(T, Cl,

.. ,

cn) = ~>k(T, Ck),


k

(9.4)

4.

3.

2.

1.

Polyisobutylene
(at 20C)
Polyethylene
(at 140C)
Polystyrene
(at 100C)
Water
(at 20C)

Polymer
designation

93

100-135

(-65)--(-80)

Tm
(OC)

Table 9.1 Thermophysical constants

1.0

1.020-1.040

0.815-0.825

0.830--0.930

X 10- 3
(kg m- 3 )

0.6

0.175-0.186

0.25-0.28

0.12-0.14

(Wm- 1 K- 1)

92-96

1.8-2.0

16.4

40-60

2.27-3.3

4.18

54-59

E
(kJ mole- 1 K- 1)

1.94

Cp
(kg kg- 1 K- 1)

264

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Using eqns (9.4) and (1.22), we find the stress-elastic strain relations for
partial stresses as follows:

Equations (9.4) and (9.5) show that in isothermal processes, the variation of elastic strain is related only to the entropy variation, and for
any kth submodel, the elastic modulus Ilk ~ PT, just as for the crosslinked rubbers. 16 ,23
It should be also noted that due to eqn (A.63) one can represent the
dissipative heat generation in eqn (9.3) as follows:
tr(u oe) = D + pjlT,

D= TPsI T = L tr(ukoep,k),

p jl T =

L tr(uk oli
k

k)

(9.6)

Equation (9.6) means that in the case of entropy elasticity under discussion, the dissipative heat generation, being equal to the mechanical
power tr(u oe), takes into account both the dissipative heating D arising
with flow, and entropy variations s due to the elastic deformations.
On the other hand, by formally introducing the heat capacity at the
constant Uk, C"k by the equality

T
C"k = CCk - - L tr[(ouk/oT)] ICk o(ohk/oT)I"kJ
p

(9.7)

similar to that common for elastic solids (see, e.g., Ref. 371, p. 229 and
Ref. 23), one can easily reduce eqn (9.3) to the form:
(9.8)
where D is the mechanical dissipation defined in eqn (9.6). Now the
question is: which value of heat capacity-Cck or C"k-is (even approximately) independent of elastic strains Ck and is only temperature dependent? According to the entropy elasticity assumption, it is CCk
because within the framework of this assumption the elastic moduli are
proportional to pT, i.e. Ilk~pT. This conclusion is in fairly good
agreement with the data. Nevertheless, the alternative statistical approach to rubber elasticity l05 discussed in Section 2.3.1, also predicts
that Il is proportional to pT, but without a significant contribution of
elastic strains in the macroscopic entropy. In this case, the assumption
of quasi-equilibrium immediately results in eqn (9.8) with C"k depending
approximately only on temperature.

Non-isothermal Flows of Polymeric Liquids

265

Only a crucial experiment can resolve the controversy between the


traditional entropic and alternative approaches to the high elasticity of
polymeric solids and liquids. No such experiment is known today.
Some attempts to derive the heat generation term from the rheological constitutive equations are described in Ref. 372.
9.3 TIME-TEMPERATURE SUPERPOSITION
Scaling methods are very helpful in obtaining experimental data in
fairly extended regions of variables and unifying them in so called
'master curves'. One of them, the time-temperature superposition, widely accepted in the rheology of viscoelastic liquids and solids, may yield
identical results when either time-dimensional parameters or temperature are varied during the experiment. To the authors' knowledge, this
scaling approach was first introduced in Ref. 373, rediscovered anew as
the famous WLF approach for representing the dynamic data in polymeric solids in Ref. 20 and was applied to the rheology of viscoelastic
polymeric liquids in Ref. 215. The applications of the method in rheology are fairly completely reviewed in Ref. 374.
Formally, the method consists of scaling variables with the dimensionality of stress by the high elastic modulus Ge , and variables with
dimensionality of time, by the mean relaxation time {}, where {}=Yj/G e
(see eqn (1.91)) and Yj is the Newtonian viscosity. According to Section
1.9, the experimental dependences contain, however, the information of
the whole set of moduli Ilk and viscosities Yjk' To use the scaling
approach efficiently, an additional assumption is needed. This is the
hypothesis of similarity, which assumes that for polymers in the viscoelastic state the ratios (}J{}k and IlJIlk> and in particular, {}J{} and
IlJG e , are temperature independent.
If the variations in temperature are not so high, the temperature
dependences of Newtonian viscosity Yj and high-elastic modulus Ge can
be approximated as follows:
Yj ~ '10

exp[E/R(T-1- To 1)],

where T and To are actual and reference Kelvin temperatures, Yjo is a


temperature-independent pre-exponential factor, R is the gas constant
and E is the activation energy of viscous flow; Yjo and E being assumed
to be temperature independent. The values of E for some polymers are
represented in Table 9.1.

266

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In rheological experiments, the condition (T - To)/T ~ 1 is commonly


satisfied in order to avoid thermal degradation. In this case, the dependences
O~Ooexp[ -m(T-To)],

Ge~ constant

(m=E/RT5) (9.9a)

can be used, which are more convenient for calculations.


Some typical examples of scaling of experimental results when the
method of time-temperature superposition is valid are presented
below.
(1) In steady isothermal simple shearing, the system of stresses is

represented as follows:
(i= 1, 2)

(9.10)

where 0"12 is the shear stress, 0"1 and 0"2 are the first and second
normal stress differences and y is the shear rate, and the approximation (9.9a) has been used. In dimensionless form, eqns (9.10)
are rewritten as
0" 12/G e

f*(yO, OdO, ... ,on/O),

O"i/Ge =

qJ~(yo,

OdO, ... , On/O)


(9.1Oa)

Thus the system of stresses depends only on the complex r = yO


which, in turn, depends on both shear rate and temperature. This
representation gives an opportunity to extend appreciably the
results of measurements made with a single instrument by varying
both the shear rate and temperature. Figure 9.1 shows the results
of time-temperature superposition for ordinary and plasticized
butyl rubber. 375 It manifests the dimensionless rheological data
for shear stress and the first normal stress difference in the form
of 'temperature independent' or 'master' curves. Various symbols
in the figure correspond to different temperatures. Here, instead
of scaling shear rate with a relaxation time, the reduced shear rate
y/y was used with the Newtonian viscosity Y/. Since Y/ = GeO, this
procedure is equivalent to scaling y with the mean relaxation time
oif the approximation (9.9a) is valid.
(2) The time-temperature method is common for the scaling of dynamic shear tests in elastic liquids where G' and Gil are represented as functions of wO. In Ref. 375, the equivalent procedure of
introducing the reduced frequency y/w was used instead for the
data shown in Fig. 9.1.

267

Non-isothermal Flows of Polymeric Liquids

~(3
1m

...

C"
~

at B}~P Be BC-P

2rC

Gt JE....
ff1Lo

Go'e m

80'C"
100*C.
f20'C 0
l/to-C

"

CI

"
II

4
3

7
9
lOS (i'p) to!l(CtJp) (Pa)

11

Fig. 9.1. Temperature-invariant plots of tangential stress (T12 and first normal
stress difference (T 1 in steady simple shear versus dimensionless shear rate y, in
comparison with these for G'(w) and G"(w) for two butyl rubbers.

(3) Figure 9.2 demonstrates the dependence of the force F on time t


for the simple extension of LDPE at constant strain rate K. Curve
1 was obtained from direct experiments with KO = 1.2 x 10 - 2 S - 1
and To = 125C. Then curves 2 and 3 were obtained for
T= 145C and 168 C with the shifted strain rates and time scales
as
(9.11)
Other temperature dependences for shifting factor aT are given in
Ref. 1.
Certain deviations from the principle of time-temperature superposition, even within the region of high elastic behaviour, have been found
for the simple elongation (Chapter 7) and are apparently related to the
occurrence of the long-range relaxation mechanism, for which it is
possible than E 1 =1= E.

268

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

GO

180 to(d)

i2Q

Fig. 9.2. The


time-temperature-invariant dependence of force F in the
simple elongation of polyethylene.
Curves 1-3 correspond to temperatures T= 125, 145 and 168C and strain
rates K=0.012, 0.0384 and 0.1 s-l,
respectively.

9.4 STEADY SIMPLE SHEAR WITH DISSIPATIVE HEATING 384


This section considers the idealized steady simple shear flow between
two parallel infinite flat plates with the gap width H =2h; one (say,
upper) plate moving with constant speed V and another being at rest.
Both plates are assumed to have the same constant temperature,
T = To. The common viscometric cartesian coordinate system used here
is similar to that shown in Fig. 4.1 but is shifted to the centreplane
between the plates. The structure of the kinematic matrices, strain rate e
and vorticity ro, is the same as shown in eqns (4.2) but now the shear
rate y depends on the transversal coordinate X2. In the following
illustrative examples it is assumed for certainty that y is positive, and
the simplified set of steady shear eqns (5.4), (5.5) and (5.10) are used,
with Yk = 0, s = and nk = n for all the modes. Then the total set of
rheological constitutive equations is of the form:

0"1 =0"11 -0"22

=2 L (J1k/n)f(Ck)(Ck -C; 1),

0" 12 =

2 L (J1k/n)f(Ck)

0"2

= 0"22 -0"33 =~L(J1k/n)[c~ + c;n -2- f(Ck)(Ck -Ck)]


k

(9.12)

y=du/dx 2

where 0"1 and 0"2 are the first and second normal stress differences, 0"12 is
the shear stress, Ck represents the maximum principle elastic strains in
the kth relaxation mode with elastic modulus J1k and relaxation time ()k,
and y is the velocity gradient in simple shearing. Note that according to

269

Non-isothermal Flows of Polymeric Liquids

the results of the previous section, parameter Ak is temperature independent.


Within the inertialess approximation, which is the exact case for the
steady shear flow under study, the shear stress 0"12 does not depend on
the spatial variable X2. Due to eqns (9.12), this fact immediately results
in the important equality:
(9.13)

r = e(T) dv/dx2 = constant = e lO V/(2h)

where v(xz) is the velocity distribution. Equation (9.13) holds because all
the stress components depend on Ck, and according to eqn (9.12),
Ck=Ck(Akr). Also used in eqn (9.13) is the approximate temperature
dependence (9.9a) for the first relaxation time e1 , with elO being a
constant pre-exponential factor.
Due to eqn (9.3), the equation for the temperature variations in
steady simple shear flow of a viscoelastic liquid is:

dZT

K
z
d-

Xz

+ 0"12 (y)y =0

(9.14)

where we have neglected the effects of the orientation phenomena and


temperature dependence on the heat conductivity, considering it as a
constant scalar, K. It should be noted that in the case of steady simple
shearing, eqn (9.14) holds for any simple liquid, independently of the
difference between eqns (9.3) and (9.8).
Equation (9.14) should be complemented by the equation:

dv
elO V
dxz = e(T) 2h

(9.15)

following from eqn (9.13), and the boundary conditions:


(9.16)
Equations (9.12), (9.14) and (9.15) form the complete set of thermorheological equations for steady shearing of the viscoelastic liquid,
which should be solved under boundary conditions (9.16).
Now, using the approximations (9.9a), we can consider all the partial
moduli Jik in eqn (9.12) to be temperature independent and introduce
the dimensionless variables and parameters as follows
Ak=ek/el=ekO/elO,
~=xz/h,

u=ve1o/hr

p=O"lZ(r, Ak)mrhz/(Ke 10 )

(9.17)

270

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where rand Ak are constants and r 15 is the shear rate defined in


isothermal flow. Then, using eqns (9.17) reduces eqns (9.14) and (9.15) to
the dimensionless form:
(9.18)
The boundary conditions (9.16) take the form:
UI~=-1 =0,

(9.19)

The solution of the boundary problem for dimensionless temperature


T* has the form: 376
(9.20)
where a is a constant of integration. Satisfying the first boundary
condition (9.19) yields the transcendent relation for the constant a:
(9.21)
The solution of eqn (9.21) represents a double-valued dependence a(P)
shown in Fig. 9.3, with the maximum possible value of P=0.88. The
double-value of a(P) means that two different temperature profiles can correspond to a fixed stressed state. When P>0.88, the steady non-isothermal
shearing flow does not exist. This situation is considered in the next section.

Fig. 9.3. The plot a(P) according to eqn (9.21).

271

Non-isothermal Flows of Polymeric Liquids

Substituting eqn (9.20) into the second eqn (9.18) and integrating the
latter with the use of the stick condition (9.19) at ~ = -1, gives:

u=2JaP/2[th(~JaP/2)+thJaP/2J
Satisfying the second boundary condition (9.19) at

~=

(9.22)

+ 1 yields:

r is = 2r J aP/2' thJaP/2

(9.23)

Equation (9.23) forms the dependence ris(r), because P and a are the
functions of r. Thus for the same stressed state of the liquid, there are two
temperature distributions and velocity profiles as described by eqns (9.20)
and (9.22), depending on the two choices of constant a in eqn (9.21).
Figure 9.4 represents the dependences r on r is for low-molecular PIB,
at various values of the dimensionless parameter
(9.24)

IT = 2111 mh2/KOlO

The quantity P is related to the parameter IT as


P=

0" 12

(9.25)

rIT/III

where 0"12 depends on r and numerical parameters: n, 'Yk = Ilk/Ill and


Ak = OklO 1. The following values of the parameters were chosen for the
PIB melt with a molecular weight of 5.7 x 10 3 377 in isothermal shear

10'

t0 2

10J

101{.

10 5

'i.J

Fig. 9.4. Plots of dimensionless shear rate r in non-isothermal steady shear


flows with dissipative heat generation versus r is for isothermal flows. Curves
1-3 represent different values of parameter TI = 5.47 x 10- 6 , 2.19 X 10- 5 and
4.92 x 10 - S, respectively.

272

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

experiments at 20C: N=2, n=1.6, 81=1Os, 8 2=0.13s, ,u1=21.5Pa,


,u2=150Pa, K=0.15Wm- 1 C- 1, E=3.39kJmole- 10C. For these
values of parameters, Y1 = 1, Y2 = 13.95, A1 = 1, and A2 = 1.3 X 10- 2. For
each value of the parameter II, there is the maximal possible value of
r = r m decreasing with the increase in II. The straight line in Fig. 9.4
corresponds to the isothermal regime of flow. It is seen that the lower the
value of II (e.g. the less the gap width H = 2h), the wider the range of r
where the dissipative heat generation is not important (r ~ ris).
As follows from eqns (9.20) and (9.22), at ~ = 0 the following conditions
hold:
u(O) = U /2

(9.26)

Then the solutions (9.20) and (9.22) of eqns (9.18) with conditions (9.19) at
~ = -1 and conditions (9.26) correspond to the following non-isothermal
steady shearing: the plate ~ = -1 is at rest and thermostatted and the
plate ~ = 0 moves with the speed U /2 and is heat-insulated. The nonisothermal shear flows of various non-Newtonian liquids, with an allowance for dissipative heat generation and different temperatures of the
plates, is considered in Ref. 376.

9.5 HYDRODYNAMIC THERMAL EXPLOSION


Let us consider the Couette plane flow sketched in Fig. 4.1 when the
mobile plate moves under given constant shear stress 0"12. As shown in
Fig. 9.4, the steady shearing between thermostatted plates is possible
only if P < P *' P * ~ 0.88, where two regimes of deformations exist.
According to Ref. 376, these correspond to steady and unsteady flows
for smaller and larger values of parameter a, respectively. In the supercritical regime (P > 0.88), the dissipative heat fails to be removed
through the wall. Therefore the time taken to achieve the steady-state
regime increases unboundedly and a steady flow is not attainable. This
is the phenomenon of hydrodynamic thermal explosion which is similar
to the thermal explosion in the case of symmetrical ignition of a flat
layer. 378
We shall investigate the unsteady flow caused by the thermal explosion
in a simplified manner. Let the characteristic time of inertia, tin = 4ph 2 /r1,
be much smaller than the relaxation characteristic time, 8=y//G e , which
in turn, is appreciably smaller than the characteristic time of the temperature variation, t T =4pCh 2 /K. Here 2h is the gap width, Y/ is the viscosity,

273

Non-isothermal Flows of Polymeric Liquids

G" is the elastic modulus, p is the density, K is the thermal conductivity


and C is the heat capacity. Because of these assumptions we can consider
the velocity profile as quasi-stationary and adjusted to the temperature
variations. Thus the only rheological characteristic of the liquid we need,
is the flow curve 0"12 (8y) with the relaxation time 8 depending on
temperature. Though the flow curve for the molten polymers is usually
well approximated by the 'power-law', the results of calculations described here are not significantly different from the results obtained in
Ref. 379, where the Newtonian rheology was employed. Therefore only
the Newtonian case is used below as an illustration.
Following Ref. 379, we consider the equation of thermal balance in the
form:
(9.27)

Cp dT/dt=q(T)-v(T - To)j(2h)

where q(T) is the dissipative heat generation, T is the liquid's temperature, To is the ambient temperature, t is the time, v is the convective heat
transfer coefficient, and 2h is the gap width between plates. Equation
(9.27) stands for the heat balance equation averaged over the gap, i.e. it is
obtained by integrating eqn (9.3) over the gap. It is assumed that
(9.28)
where v depends on the parameters of flow of the liquid and its physical
characteristics. When the plates are thermostatted, T --+ To and v--+ 00.
Note that eqn (9.28) is strictly true only when the temperature is
distributed almost uniformly across the gap, i.e. when the following
condition for the Biot criterion,
Bi=vh/K~

(9.29)

is satisfied. When it is approximately the case, eqn (9.27) can be considered as the equation of thermal conductivity averaged over the gap.
Using this equation below, we first need to find the type of heat source
q(T) to analyse the behaviour of the solution. As mentioned proviously,
we consider only the simplest case of a Newtonian liquid that does not
have any effect on our qualitative analysis. In this approximation,

q(T)=~f: 0"12ydx2~0"12Y
and the right-hand side of eqn (9.30) does not depend on
the time-dependent temperature T(t).

(9.30)
X2

but only on

274

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

If the dependence of viscosity on temperature is assumed to be of the


form shown in eqn (9.9a), eqn (9.27) of thermal balance can be represented
in the dimensionless form:

d0/dr=exp(0)-0/K

(9.31 )

where
q(To)E

r= CpRT6t,

K = -.:2q:.. :. T....::o-,-=)h_E
(
vRT6

(9.32)

and in our rough approximation the parameter v is assumed to be


constant. The plus sign in the exponent (eqn (9.31)) corresponds to
shearing under given constant shear stress, 0"12 = constant, and the minus
sign to shearing under a given constant velocity V of the mobile plate.
Equation (9.31) shows that in the case V = constant, a steady shear flow
is always attainable with any value of parameter K. For the case
0" 12 = constant, there are two very different regimes of start-up flow. If
K < K*( = l/e), the steady shear flow eventually occurs, accompanied by a
relatively small dissipative heat generation. If however, K > K*, the
progressive dissipative heating in flow leads to a monotonic increase in
d0/dr which, in turn, results in a very large increase in temperature. This
is the phenomenon of hydrodynamic thermal explosion. The physical
reason for this is the process of auto-acceleration which can happen
without constraint on the velocity field: an increase in temperature due to
the dissipative heating diminishes the viscosity, which, in turn, accelerates
the flow and results in the increase of dissipative heat.
The conditions for the onset of hydrodynamic thermal explosion were
theoretically predicted in Refs 380-382 for the Poiseuille flow in a long
tube under a given pressure drop. The phenomenon was observed
experimentally383 in a Couette flow of a Newtonian fluid between coaxial
cylinders.

9.6 STEADY SHEARING FLOW OF POLYMERIC LIQUIDS


BETWEEN ROTATING DISKS WITH AN ACCOUNT OF
DISSIPATIVE HEATING
The sealing effect of the rotating shaft sealing377.384 is based on the
Weissenberg effect created by the intense shearing of a polymer liquid
between rotating disks, the working units of the device. Therefore the
long-term behaviour of normal stresses under the action of dissipative

Non-isothermal Flows of Polymeric Liquids

275

heat is very important here. A brief description of the device is given in


Section 9.7.
9.6.1 The Scheme of Apparatus
This apparatus was purpose-built to study the shearing flow of polymeric
fluids between rotating disks, while simultaneously measuring the radial
normal stresses. The design of the experimental device is sketched in Fig.
9.5. The polymer (1) is in a gap of width H = 2h, between the disks (2 and
3) of radius R 1 . The lower disk (2) rotates with a constant angular speed
n, while the upper one (3) is at rest and rigidly connected to the
cylindrical body (4). The disk (3) has a circular hole of radius R 2 In the
experiments, Rl =40 mm, and the radius of the hole was varied: R2 = 1, 7
and 15 mm. Part of the cylinder above the disk (3) was partially filled by
the same polymer (5), which was practically non-deformed. This device
served as a buffer to separate the water (6) that filled up to the rest of the
body, from the rotating shaft. It should be noted that in the presence of
water, the polymer sometimes broke away from the walls. When the
mobile disk (2) is rotated, normal stresses are developed in the polymer
layer between two disks. These were transferred through the hole and
through the stationary layer (6) of the polymer to the water, creating
pressure Pl in the water layer. This pressure is measured by a manometer
(7). The pressure P2 at the periphery of the rotating disk is approximately
equal to zero.
To keep the temperatures of the stationary disk (3) and body (4)
constant, they were made hollow with a coolant circulating through

5
/;
1

L-1~2S~~~~Fl=:-- z

Fig. 9.5. Sketch of experimental device


for measuring normal radial stresses: 1,
shearing polymer layer; 2, rotating disk;
3, fixed disk; 4, cylindrical body; 5,
fixed polymer layer; 6, water layer; 7,
manometer.

276

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

them. The mobile disk (2) was thermos tatted with a free jet of a liquid
coolant or was made heat-proof.
It was found that when R2 ~ h ~ R 1 , the effect of the hole on the flow of
polymer between the disk was negligible. When R2 was greater than h
(say, R2=7 and 15mm, h=2mm), measures were undertaken to exclude
the contribution of central region to the measured pressure Pl' To do
this, the experiments were conducted both with the gap (5) completely
filled up with the polymer, and with only the central or marginal zone
filled. The effects of rotation speed n and geometrical parameters hand
R2 on the value of pressure Pl and the torque M were measured. The
studies were carried out with PIB of molecular mass M ~ 5.7 X 103 , whose
Newtonian viscosity 1'/ at 20C was: 1'/ ~ 103 Pa s - 1.
9.6.2 Theoretical Analysis
It is convenient to study the flow between the disks in the cylindrical

coordinate system {r, cp, z} with the origin placed in the middle plane
between the disks. One disk is at rest and another rotates with a constant
angular speed n. The equation for the steady temperature variations,
with allowance for the angular symmetry, is of the form:
K[T,zz + r -1 (rTor)" ] =

tr(O' e)

(9.33)

where T is the temperature, K is the thermal conductivity, 0' and e are


stress and strain rate tensors, and the lower indices after a comma mean
the space partial derivatives. The boundary conditions for temperature
are of the form:
(9.34)
The boundary conditions for the velocity are:

vl z = -h =0,

(9.35)

The boundary conditions for the stress are discussed later. Note that the
last boundary condition (9.34) can also be used to describe the temperature distribution when the radius of hole R2 (Fig. 9.5) is small enough and
the effect of hole on the polymer flow can be neglected.
The assumption h ~ Rl results in the inequality:
(9.36)
which in turn, reduces eqn (9.33) to the simplified form:
KT,zz =

tr(O' e)

(9.37)

Non-isothermal Flows of Polymeric Liquids

277

For this equation, the boundary conditions (9.34) at r=O and r=R 1
should be omitted. Then at each fixed value of r, the solution of this
problem coincides, up to isotropic pressure, with that valid for the flow
between parallel plates, considered in the previous section. Therefore, the
'rheological' stress components, 0"1> 0"2 and 0"12, and the dimensionless
shear rate r are some single-valued functions of the parameter
(9.38)

i.e. of radius r. These are estimated by formulae given in Section 6.4.


Evidently, ineqn (9.36) and eqn (9.37) are violated at the centre and at the
rims of the disks. In these regions, the temperature fields and the values of
shear rates r should be estimated separately.
In the central zone where the shear rate is negligible, we can also
neglect the dissipative heating to obtain a good approximation, that in
this zone, T:::::; To. More tedious calculations are needed to obtain the
radial temperature and shear rate distributions averaged over the gap.
An asymptotic solution of the problem near the external rim of the
rotating disk, r = R 2 , was obtained and matched with the solution in the
main region (R1 <r<R2). For details of the calculations, see Refs 377
and 384.
With a known radial distribution of shear rate r, one can use formulae
(9.12) to calculate the rheological components of the stress tensor. Finally,
integrating the equation of stress equilibrium

with approximate conditions:

yields:
P1=P2+

+0"2
---dr

R 2 0"1

Rl

(9.39)

where 0" 1 and 0"2 are the first and second normal stress differences defined
in eqn (9.12). The expression for the torque on the rotating disk is:
(9.40)

where the shear stress 0"12 is also defined in eqn (9.12).

278

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

9.6.3 Comparison of Theoretical Calculations with Experiments


As mentioned above, the studies were carried out with PIB of molecular
mass M~5.7 x 10 3 , whose Newtonian viscosity 1J at 20C was:
1J ~ 10 3 Pa S-l. For this PIB, thermophysical and rheological parameters
in eqn (9.12) have been discussed in Section 9.4. These were found at
temperature T = 20C using the flow curve and the first normal stress
difference plot. All the calculations were made under the assumption that

pz=O.
Figure 9.6 shows the plots of the pressure P1 (Fig. 9.6(a)) and torque M
(Fig. 9.6(b)) versus the speed of revolution of the disk, n. The curves and
symbols 1, 2 and 3 correspond to the radii R z of the hole when R2 = 1, 7
and 15 mm, respectively, when one of the disks was thermostatted at
temperature To = 20C and another was thermo-insulated. The curves
and symbols 4 correspond to the case R z = 7 mm with both disks being
thermos tatted. The external rims of the disks (R1 = 40 mm) were thermostatted in both cases.
The symbols denote the results obtained by measuring the pressure P1
by a manometer. The solid lines represent the results of calculations
which ignored the thermostatting effect of the disk's rims, and the
dashed lines indicate an allowance for this effect. It is seen that in these
experiments, thermostatting the rims had almost no effect on the results
of the calculations. For more viscous liquids, however, where under the
same conditions, the effects of dissipative heat generation are more
significant, there is a noticeable increase in the contribution of the
stresses to the pressure P1 and torque M due to the motions near the
nm.
Figure 9.6 shows that at fixed value of n, the larger the radius R z of an
internal hole, the less is the pressure P1 and torque M. At high values of
R z , the dissipative heating manifests itself relatively strongly, with the
maxima on the plots. The maxima are related to the dependence qris )
shown in Fig. 9.4. In the central zone where the dissipative heating is not
so important, qris ) increases, whereas at the disk's periphery where the
heating is essential, it decreases. When the value of R z is high enough, the
increase in the relative contribution of the rim region results in the
appearance of the maxima.
At a fixed geometry and with high values of n, the pressure P1 and
torque M are, obviously, larger for both thermostatting disks than for
one being thermo-insulated (see symbols 2 and 4 in Fig. 9.6). The
coincidence of the dependences 2 and 4 for small rotation speeds indicates
that the flow is almost isothermal.

279

Non-isothermal Flows of Polymeric Liquids

P(MPo.J
0.24
()

1
---------0.16

e-1

0.08

0-2
v-3

()-4-

v
-----

0r-------------------------________

(Nm)2.0

('0

80

160

2M

n(RPM)

Fig. 9.6. Plots of (a) pressure P and (b) torque M versus the number of
revolutions Ii of the rotating disk in a self-heated flow. The radius of disks
R 1 = 40 mm, and the distance between discs H = 1.7 mm. Curves 1-3 correspond
to the radius of inner hole R2 = 1, 7 and 15 mm, respectively, with one disk being
thermostatted; curve 4 corresponds to R2 = 7 mm with both disks being thermostatted. Points represent experimental data; solid lines are the results of calculations with thermostatting effect of the disk's rims ignored; dashed lines are those
with allowance for this effect.

The calculations also showed that the pressure and the torque at fixed
values of rotation speed and radii Rl and R 2 , sharply decrease with an
increase in the gap width H or with growing temperature on the
thermostatted surfaces.

280

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Fig. 9.7. Calculated plots of maximum


( - - ) and averaged over the gap (- - --)
temperature differences ~ T = T - To versus
radius in the self-heated shearing flow
of a polymer between rotating disks;
n= 100 rpm. One disk is heat-insulated, another is thermostatted. Curves 1 and 2
correspond to H = 2 and 4 mm.

Figure 9.7 displays the calculated radial distributions of temperature


averaged over the gap, <T) (dashed lines), and the maximum temperature
in the gap, Tm (solid lines), for two values of gap width. In these
calculations, only one disk was thermostatted. According to the calculations, the maximum of the averaged temperature occurs near the thermostatted rim.

9.7 SEALING OF A ROTATING SHAFT WITH


POLYMERIC LIQUIDS
Figure 9.8(a) illustrates one of the designs38s.386 of sealing 387 based on
the normal stress effect. A polymeric liquid (5) is between a rotating disk
(2) of external radius Rb connecting to the rotating shaft (1) and the
stationary disk (4) connected to the cylinder body (3). The pressure
difference /1p = Pi - P2 occurs in this device because of the normal stresses
arising in the sheared polymeric liquid in the gap between the disks, and
directed along the radius to the rotating shaft. The sleeve (6) of external
radius R2 which can slide along the shaft and the stationary disk fixed to
the body, prevents spinning up to the polymeric fluid on to the shaft.
There are also other design features to prevent this effect.
In order to increase the effect of normal stresses, such high molecular
and large viscosity fluids as PIB P-20 were used. With the Newtonian
viscosity 1'/ as large as 106 Pa s - 1 (at 20 Qq, the high dissipative heating
arising in the intense shearing of the fluids makes the effects of nonisothermality very important. To stabilize the work of the sealing, one or
both disks were thermostatted.

Non-isothermal Flows of Polymeric Liquids

281

Fig. 9.8. Sketches (a) single and (b) connected in series sealing units with
polymeric liquid for the rotary shaft: 1, rotating shaft; 2, rotating disk; 3,
cylindrical body; 4, stationary disk; 5, sheared polymeric liquid; 6, sleeve.

Estimations of the pressure drop and torque in the sealing are carried
out on the basis of the approach developed in Section 9.6 for the
non-isothermal shearing of a viscoelastic liquid between rotating disks
with holes in their centres.
The stability of functioning the shaft sealing when the pressure difference changes is based on the change in the polymer volume in the
working gap between the disks.
It should be also noted that in the case where the upper region, which
is under the pressure Pi, is filled up with a fluid such as water, and the
disks are made of metal, the polymer can break away from the wall
with the formation of a thin water film. This apparently occurs due to
surface tension forces. In this case, coating the disks with a hard
polymer prevents the breaking away of polymer from the walls. When
there is no motion, or the water is replaced by a gas, this phenomenon
is not observed. These phenomena are discussed in more detail in
Chapter 12.
The single sealing can maintain a not so high pressure difference:
Llp~0.1-O.3 MPa at n~ 100 rpm is the usual order of magnitudes. To
increase the total pressure drop in the sealing, these can be connected in
series, as shown in Fig. 9.8(b). If each sealing element can maintain the
pressure Llp., the total pressure difference in the 'battery' of N elements is
Llp = N Llps. The intermediate regions in the sealing battery are usually

282

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

filled up with a gas, the pressure of which is maintained by a controlling


device.
In conclusion, it should be noted that the polymeric fluids can also be
used as a working body in thrust bearings,388 journal bearings 389 and
shaft centring devices.
9.8 NON-ISOTHERMAL CAPILLARY FLOWS WITH AN
ACCOUNT OF DISSIPATIVE HEAT GENERATION

This section focuses on general fluid mechanical problems whose qualitative peculiarities are almost independent of the rheology of liquids.
Nevertheless, the problems discussed in this section can hopefully be
useful for the modelling of non-isothermal high De flows of polymeric
fluids in various problems of polymer processing.
9.8.1 The Problem Formulation

Let us consider the laminar and non-isothermal flow of an incompressible


liquid in a long circular pipe of radius R and length L (L ~ R). A similar
isothermal problem for viscoelastic liquid is considered in Chapter 8. As
in Section 9.4, we assume that
(9.41)

where tin, f) and t y are the characteristic times of inertia, relaxation and
temperature variations, respectively. When ineqs (9.41) are satisfied, the
non-isothermal steady flows of viscoelastic liquids in very long tubes are
similar to those of the inelastic power liquids, which in turn, do not differ
qualitatively from the similar flows of highly viscous Newtonian liquids.
The analysis of the problem 390 showed that the thermal entrance
effects with the dissipative nonlinearities taken into account, play an
important role in the formation of the flow. These effects do not lead to
the hydrodynamic thermal explosion phenomena 38o but to a possible
drastic jump-like transition, with a continuous change in the pressure
drop, from a low-temperature steady flow to a high-temperature one, and
back. Moreover, hysteresis occurs in the forth and back transitions when
the critical conditions for these do not coincide. It was noticed in Ref. 390
that these effects belong to the same class of phenomena as ignition and
extinction in combustion theory378 and in magnetic gas dynamics,391 due
to the nonlinearities in the dependences of the reaction rate and electric
conduction on temperature.

283

Non-isothermal Flows of Polymeric Liquids

To simplify the problem, let us assume that the temperature distribution in the tube cross-section is almost constant and that the conductive
longitudinal heat flux is negligible in comparison with that in the
cross-sectional direction and with axial convective heat flux. As mentioned previously, the first assumption is appropriate for small Biot
numbers, Bi = VR/K ~ 1, and the second one with large Peclet numbers,
Pe = pCQ/nRK ~ 1, where v, K and C are the coefficients of crosssectional heat transfer, thermal conductivity and heat capacity, respectively, p is the density and Q is the flow rate. If the condition Bi ~ 1 is
not satisfied and cross-sectional temperature variations are essential,
the approach given below is still qualitatively valid when operated with
the temperature averaged over cross-section and effective heat transfer
coefficient v.
With these assumptions, the equation of steady thermal balance per
unit liquid volume is written in the form:
pCQ dT
2v
-=q(T)--(T-To)
nR2 dz
R

(9.42)

where T and To are the temperatures of fluid and environment, q(T) is


the dissipative heat, v is defined by eqn (9.28) and z is the longitudinal
coordinate.
When the dissipative heat is negligible (q(T) ~O), eqn (9.42) can serve as
a good approximation to describe the processes of heating or cooling a
viscous liquid moving along a tube. 392 ,393 To determine the dissipative
heat source q(T) in the general case, one has to consider momentum
balance and constitutive equations for a liquid, complementary to eqn
(9.42).
In order to find the expression for the function q(T) one should employ
the common procedure considering the flow along the tube as quasiPoiseuille, with the same distributions of shear stress and velocity across
the tube as in isothermal flow and 'slow' variations of all profiles along
the tube. 390 We use further the Newtonian rheology: O"rz= I'/(T) dv/dr,
with the approximation for temperature dependence of viscosity:
I'/(T)

~I'/o

exp[ -meT - To)]

(1'/0, m = constant)

(9.43)

Using the definitions of flow rate Q and dissipative heat q(T):


Q=2n

f:

vrdr,

2
q(T)=R2

fR (r-dP)
dv
- -rdr
0

2 dz

dr

(9.44)

284

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

we can obtain the expressions for the pressure gradient along the tube
and dissipative source as follows:
8t/oQ

-dp/dz~--4

nR

exp[ -m(T - To)]

(9.45)
(9.46)

Note that unlike the isothermal Poiseuille flow, in the non-isothermal


case, the pressure distribution along the pipe is nonlinear because of the
longitudinal variations of temperature.
Substituting eqn (9.46) into eqn (9.42) reduces the thermal balance
equation to the form:

Further, we consider the simplest case when the temperature of the


liquid at the tube's entry and of the environment are equal, i.e.
(9.48)
Integrating eqn (9.45) along the whole tube gives the expression for the
pressure drop, J1p == p(O) - p(L), as follows:
8t/oQ
J1p = --4
nR

fL
0

(9.49)

exp[ - m( T - To)] dz

Two regimes of steady pipe flow are usually under study: ratecontrolled (flow rate Q is given) and pressure-controlled (pressure drop
J1p is given). For the non-isothermal flow under study, the first case is
related to the solution of the initial problem (9.47), (9.48) with given
parameter Q. The second problem is related to the solution of that with
the additional nonlinear integral condition (9.49) serving to determine
the flow rate Q.
By introducing the dimensionless variables and parameters
0=m(T-To),

m
P=-J1p
pC

'

(9.50)

Non-isothermal Flows of Polymeric Liquids

285

we can reduce the set of eqns (9.45), (9.47) and (9.49) with the initial
condition (9.48) to the following dimensionless form:

w d8/d~ = w2 exp( - 8) - B8,


P=w

exp( -8) d~

8(0) = 0,
(9.51)

For the sake of brevity, the notation for the dimensionless variables is
retained below, i.e. w is retained as the flow rate, P as the pressure drop,
8 as the temperature, and B as the heat transfer coefficient.
9.8.2 Some Results of a Numerical Study
Let us consider first the results of the numerical solution of the problem.
Of primary interest here is the dependence of the temperature at the tube
exit, 8 1 == 8(1), on the given pressure drop P. A family (1-7) of such
curves, corresponding to different values of heat transfer coefficient B, is
depicted in Fig. 9.9. The straight line (1) in the figure describes the
adiabatic flow when B=O and has the simplest form: 8 1 =P. The result
immediately follows from eqns (9.51). As shown in Ref. 394, at w -+ 00
there are the asymptotic expressions:
(9.52)
independently of the value of parameter B. This means that the adiabatic
regime represented by the straight line (1) in Fig. 9.9 is the asymptote for
all the temperature distributions.
When the value of parameter B is small enough, e.g. B= 10 and 30 as
for the curves 2 and 3 in Fig. 9.9, the exit temperature 8 1 increases
monotonically with the pressure drop P increasing. Thus, only one steady
temperature distribution is possible in this case. This is the so-called
low-temperature regime which corresponds physically to the slow flow of
01
84

Fig. 9.9. Plots of relative exit temperature 8 1 versus assigned pressure drop P
at various values of heat transfer coefficient B. Curves 1-7 represent B = 0, 10,
30, 60, 100, 250 and 500, respectively.

286

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

a low viscosity liquid. In this case, I8 I ~ 1 and one can neglect the effect
of temperature on the dissipative heat generation:
q(8) == w 2 exp( - 8) ~ w 2 = constant

(9.53)

With approximation (9.53), the solution of the problem (9.51) has the
form:
w~P

(9.54)

In the high-temperature regime where 8 ~ 1, the adiabatic solution (9.52)


gives a good approximation.
With the value of parameter B increasing, the curves in Fig. 9.9 start to
increase more sharply. When the value of B reaches a certain critical
value, B* ~ 60 (P = P * ~ 5.65), the corresponding curve (curve 4) ascends
almost vertically.
With a further increase in the B value, the curves 5, 6 and 7 (B = 100,
250 and 500) in Fig. 9.9 take a characteristic S-shaped form. For the
sake of discussion, let us select curve 6 with the value B = 250. Three
branches can be distinguished on the curve: the lower or low-temperature branch {0~P~P+,0~81~8+}, the intermediate one, and the
upper or high-temperature branch {P):P-,81):8_} corresponding to
an almost adiabatic regime. It is important to note that both the lowand high-temperature branches are increasing (d8ddP>0), whereas the
intermediate one is decreasing (d8ddP < 0). When moving along the
curve from the region of low temperatures, gradually increasing the
pressure drop P, the steady thermal regime is suddenly changed in a
jump-wise manner after reaching the point P +, from the value 8+ to a
very high value (not shown in Fig. 9.9) on the high-temperature
adiabatic branch. The reverse transition with pressure drop P decreasing, also occurs after reaching the value P _, with a jump transition in
temperature, from the higher value 8 _ on the adiabatic branch to a
certain value on the low-temperature branch. The forth and back
temperature transitions form hysteresis, very similar to the ignition and
extinction phenomena in the thermal combustion theory.378 Therefore
for the non-isothermal flows under study one can speak about hydrodynamic thermal ignition and extinction, the critical conditions for
which, as in the theory of combustion, do not coincide. As seen from
Fig. 9.9, the hysteresis and critical conditions are possible only when
B>B* (~60). When B<B* the steady temperature 81 increases monotonically with a growing pressure drop P; such flow situations are
known as 'crisis-less flows'.

Non-isothermal Flows of Polymeric Liquids

287

The phenomena discussed above take place when the constant


pressure drop P is given. In the case when the constant flow rate w
is given, the dependences 0 1 on In ware monotonically increased
and saturated with an adiabatic temperature 0~ when In w ~ 00. The
normalized plots 0d0~ versus In w presented in Fig. 9.10 show that
there is only one steady non-isothermal flow regime in this case. These
regularities are also illustrated in Fig. 9.11, which depicts the plots 1-6
of pressure drop P versus the logarithm of flow rate, In w, at various
values of parameter B. When values of B are small enough (curves
1-3), there is only one value of w corresponding to a given value
of P. With a sufficiently large value of B (curves 4-6), there are three
steady values of flow rate w corresponding to a given value of
pressure drop P from the interval P _ < P < P +. These correspond to
the low, intermediate and high temperature regimes; the intermediate
regime being unstable and relating to an unsteady solution. With a
gradually changing pressure drop P, the transition from small values of
flow rate to a large one and back is drastic and takes place with a
hysteresis.

"

..:In"
f}

DO1

0.8
0.1i

t5

0.2

'I

~ ,g/UI1

Fig_ 9.10. Plots of relative exit temperature e/e o versus assigned flow rate
w at various values of heat transfer
coefficient B. Curves 1-7 represent
B=O, 10, 30, 60, 100, 250 and 500,
respectively.

2 3
5" 6 7tnJ
Fig. 9.11. Plots of pressure drop P
versus flow rate w for various values of
heat transfer coefficient B. Curves 1-7
represent B=O, 10,30,60,100,250 and
500, respectively.

9.8.3 An Analytical Approximation

To analyse ignition and extinction phenomena in combustion theory


as well as the hysteresis effects for simple magnetic gas dynamic
flows, a simplified analytical approach called the 'zero-dimensional
method' was elaborated and successfully applied to the solution of these

288

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

problems. 391 ,395 This method was also applied to the analysis of nonisothermal flow in long pipes. 39o Instead of a temperature distribution
along the tube, this method operates with a certain mean value of
temperature 0. To simplify the notations, hereafter the circumflex is
omitted. The key element in this method is using the finite-differential
approximation of the derivative dE>/d~ in the simplified 'zero-dimensional' form: dE>/d~~E>. Then eqns (9.51) are reduced to the simplified
algebraic form:
wE>=w 2 exp( -E-BE>,

(9.55)

P=wexpCE

When the flow is under given pressure drop P, the relations can be
rewritten in the following form:
(9.56)
Thus the study of the steady regimes in this approach is related to the
easy analysis of the intersection of the straight line q2(E with the curve
ql(E, the left-hand side of eqn (9.56). The geometrical interpretation of
this analysis is presented in Fig. 9.12. Figure 9.12(a) shows the lowtemperature regime. Figure 9.12(b) demonstrates the critical conditions
for the transition from the low- to the high-temperature regime and the
occurrence of a single contact point between the line ql and curve
Q2' Figure 9.12(c) illustrates the phenomena of hydrodynamic thermal
c)

Fig. 9.12.

Determining the steady points of eqn (9.56): (a), (b) and (c) correspond
to various values of parameters Band P (see text).

ignition (contact line 2) and extinction (contact line 4), with the jump-wise
transition from low- to high-temperature regions and the occurrence of
the hysteresis.
Figure 9.13 compares the results of exact calculations (solid lines) and
approximate ones (dashed lines), represented as the plots of the exit

Non-isothermal Flows of Polymeric Liquids

289

temperatures e versus the pressure drop P. Curves 1 and 2 correspond to


the values of B = 30 and 100, respectively. The respective curves practically coincide at small values of e, including the critical point of ignition.
However, in the vicinity of extinction point the discrepancy between these
is significant.
Based on eqn (9.56), easy calculations were performed 390 to find the
dependences of critical values for the heat transfer coefficients B + and B_

Fig. 9.13. Comparison between exact


( - - ) and approximate (- - -- -) calculations of plots 8 versus P for two values of
parameter B: 1, B = 30; 2, B = 100.

and temperatures e + and e _ on the pressure drop P, where indices


+( -) correspond to ignition (extinction). The corresponding plots presented in Fig. 9.14 show that the whole region of the parameters is
divided into two zones. The hysteresis zone 1 is confined between lower
and upper branches corresponding to the critical conditions of ignition and
extinction, respectively. Zone 2 is thermally crisisless. The dashed and
solid lines in Fig. 9.14 show the results of approximate and exact
calculations, repectively, and the shaded area represents the discrepancies

Fig. 9.14. Plots of critical parameters B and 8 versus pressure drop P.


Solid and dashed lines represent exact and approximate calculations,
respectively.

290

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

between them. As seen, the approximate calculations describe well


enough the critical temperature parameters.
The approximate calculations allow us also to find easily the values of
parameters B*, P *' 0* and w* corresponding to the beginning of the
hysteresis zone. These are defined in Fig. 9.l4(a) and (b) as the intersection
points of the upper and lower branches. The values of the parameters are:
P*=4,

w* =B* =4 exp(2)~30

(9.57)

To illustrate the onset of thermal hysteresis phenomena, consider the


example of a flow of castor oil in a long capillary tube. In the temperature
interval 9-40C, the parameters of the liquid are: C = 2.14 J g - lO C-1,
p=O.964gcm- 3 , m=O.085C- 1 . Let To=9C, R=O.1 cm, L=20cm.
Then the dimensionless critical conditions (9.57) give the respective
dimensional values: AT*=23.5C, Ap*~100MPa, Q*=375cm 3 s-1,
v*=6.16Jcm- 2 s-1, and Re=150. Thus all the conditions for onset of
the critical thermal flow phenomena are satisfied with the parameters of
flow looking realistic enough. In order to lower the critical value of the
pressure drop, Ap*, one can use a preliminary heating of the liquid so that
the entry temperature, Tlz=o, may be higher than the ambient temperature To.
9.8.4 Flows in Long Tubes
It is also of interest to analyse the limiting case of the 'infinitely long'
tube, when B ~ 00 and P ~ 00. Then two 'zero-dimensional' equations
(9.55) are reduced to the single one-parametric equation:

exp(0)-K0=O,

(9.58)

where the parameter K is the ratio of heat transfer to heat generation. 380
In this case, the high-temperature steady flow along with the extinction
does not exist. Instead, the phenomenon of hydrodynamic thermal
explosion occurs here with the obvious critical ignition condition: 380

9.9 NON-ISOTHERMAL ELONGATIONAL DEFORMATION


OF POLYMERIC LIQUIDS
As mentioned previously, the non-isothermal effects in simple elongational flows of highly viscous polymeric liquids violate, strictly speaking,

Non-isothermal Flows of Polymeric Liquids

291

the one-dimensional elongational flow kinematics commonly accepted for


isothermal flows. The main reason for this is the heat conductivity
phenomena starting from the surface of the specimen and propagating in
a cross-sectional direction, orthogonal to the direction of extension. This
leads to the cross-sectional distribution of temperature, which in turn,
immediately initiates the occurrence of tangential stresses. If, however, a
stretched specimen is very thin, the characteristic time t y of the thermal
process can be considerably less than the characteristic relaxation time e.
Thus in the case when
(9.59)
where R is the characteristic cross-sectional size of the sample and r:x is the
thermal diffusivity coefficient, the assumption of a quick thermal adaptation can be true. Along with the negligible effects of dissipative heating
which usually hold because of slowness of elongation flows, this assumption enables us, at least in some cases, to consider still the simplified
one-dimensional description of non-isothermal elongation flows. Below,
we consider two cases: when the assumption (9.59) of quick thermal
adaptation is seemingly true, and the example when it is violated.
9.9.1 An Example of Quick Thermal Adaptation

In Ref. 396 the calculations according to eqns (6.16) with nk=l, 8=0 and
e~ = e(T) were performed for non-isothermal elongation flows of polymers. The calculations were compared with elongational experiments298.397 for a polystyrene under a constant cooling rate and under
sudden changes in temperature. These experiments were carried out
under a constant strain rate and under a constant pulling rate.
The relative temperature shift factor
I ogaT=

-C1C2(T-Tr)
(C 2 + Tr- Tg)(C 2 + T - Tg)

(9.60)

was employed in the calculations. Here T, Tr and Tg are actual,


reference and glass transition temperatures, respectively, with the
values: 398 Tr =160C, T g =94C, C 1 =15.5C and C 2 =45.5C. The
increased value of Tg = 108 DC was also employed under cooling with
a constant cooling rate, following the recommendations of Ref. 398.
Six nonlinear Maxwellian modes were used in these calculations.
The parameters of the model, ek and 11k> were determined in the
linear region of deformations to describe quite accurately the plot
G"(w) for polystyrene TC-30 (Shell) at 160C 399 within six orders of

292

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

frequency variations. Table 9.2 demonstrates the values of the parameters


and Fig. 9.15 shows the dependence GI/(w). With these data, the
calculations predicted well enough the data obtained by different authors
for isothermal elongational flows under a given pulling rate and a given
strain rate.
Figure 9.16 displays the comparison between the calculations and
data 396 for elongation flows with an initial temperature of 160C and a
strain rate equal to 0.0528 s -1 for two different cooling rates. R1 = 0.91
Table 9.2
Model parameters of a commercial polystyrene TC 3-30 Shell at
160 DC using six modes; s = 1.0 x
10 - 5 (Table 1 in Ref. 396)
11k

(Pas- 1 )

2.96 x
8.49 X
1.48 X
1.78 X
3.12 X
8.76 X

(}k(S-l)

10 5
104
104

3.01 X
3.11 x
3.54 X
3.22 X
2.38 X
1.64 X

10 3
10 2
10 1

10 1
10
10- 1
10- 2
10- 3
10- 4

6
./"

......
0

....'

Q..

"=

<!>

Cl

...J

2
-3

-2

-I

Log W (5-')

Fig. 9.15. Frequency dependence on shear loss modulus G"(w). The solid line
shows the result of linear viscoelastic fitting with the values of parameters given in
Table 9.2. Points correspond to the data for polystyrene TC 3-30 (Shell) at
160 C. 396
D

Non-isothermal Flows of Polymeric Liquids

293

9.0 r - - - - - - - - r - - - - - - . - - - - - - - ,

8.0

70

t:II

...J 6.0

5.0

4.01--_ _ _ _---'-_ _ _ _ _- ! -_ _ _ _----'


-1.0
2.0
o
1.0
Log (5)

Fig. 9.16. Time dependence of elongational viscosity l1e for two different cooling
rates R with an initial temperature of 160C at a strain rate of 0.0528 S-I:
(---) prediction with 'equilibrium' value Tg=94C, (----) prediction with
Tg=108C; 0, R=0.91Cs- 1 ; 6., R=2.18Cs-I.396

and R2 =2.18 C S-l. Solid lines show the predictions with Tg =94 C and
dashed lines with Tg = 108 0c. Figures 9.17 and 9.18 compare the results
of calculations 396 and data 397 for the restructuring of the tensile stress
with temperature jumps (Fig. 9.17) and drops (Fig. 9.18) in elongational
flows under a constant pulling speed.
Though there were no data available for sample dimension variations
during the experiments, the good agreement between the predictions and
the data demonstrated in Figs 9.15-9.18 indirectly testifies in favour of
the quick thermal adaptation achieved in these experiments.

294

Nonlinear Viscoelastic Effects in Flows of Polymer Melts


70 r - - - r - - - - - r - - - - - r - - - - - - - - - - ,
EO' 0.08335-1

::;.

to

CIl

60

...J

119C

50

7.0 r - - - - , - - - - - , - - - - - - - r - - - - - - - ,

10

60

CIl

5.0

_~

-----'-0-1-----::-,-::-1

'------L--l

10

Log t (5)

2.0

3.0

Fig. 9.17. Time dependence of tensile stress (J under a constant pulling speed (an
initial strain rate of 0.0833 S-I). The temperature jumps from 119 to 137C at
times indicated by arrows in (a) and (b). Solid lines are predictions with
temperature jumps, dashed lines are isothermal predictions, and points represent
experimental data.396.397

9.9.2 Non-isothermal Strain Recovery


Non-isothermal strain recovery is one of many possible examples of
viscoelastic deformation when the characteristic relaxation and thermal
times are of the same order.
Consider the case of non-isothermal retardation when, after the
isothermal extension of a viscoelastic cylindrical sample up to the length 1
and radius R, the stress is instantaneously dropped to zero. After this the
sample is allowed to contract with time at the expense of the elastic
energy accumulated in extension. An experiment of non-isothermal strain
recovery was carried out for PIB p_20197 with a maximum Newtonian
viscosity at 22C, '1 = 1.3 X 106 Pa s - 1, relaxation time e~ 400 s, and
retardation time er~80s. The thermal diffusivity K~7xl0-4cm2s-1.

295

Non-isothermal Flows of Polymeric Liquids


7.0 , - - - - - , - - - - - - , . - - - - - - - - ,

119"C /
,/

137'C' 0

5.0

7.0,-----,-----.--------,

"ti'

.e;,
<0 60

50

/
0

//

I19"C//

L,
1

C-...-/--'-------'---_

1.0

Log t (s)

2.0

3.0

Fig. 9.18. Time dependence of tensile stress (J under a constant pulling speed (an
initial strain rate of 0.0833 s -1). The temperature drops from 137 to 119C at
times indicated by arrows in (a) and (b). Solid lines are predictions with
temperature jumps, dashed lines are isothermal predictions, and points represent
experimental data. 396 ,397

Cylindrical specimens have been isothermally pre-extended at temperatures of 22 and 70C. The processes of strain recovery were carried out
both isothermally at the extension temperatures, and non-isothermally at
free cooling with an environmental temperature of 22C after extension
at 70 0C. The characteristic thermal time tT ~ 100 s, was of the same order
as the retardation time.
Figure 9.19 shows the time dependences for the ratio 1./1 in isothermal
and non-isothermal regimes of retardation. Here f.(t) is the length of the
contracting specimen. In all the three cases of retardation, the samples
were stretched up to the stress level (J = 1.3 X 10 5 Pa with elastic strain
Ii = 2.6. As shown in Chapter 4 (see also Ref. 287), the extension regime
with which certain values of (J and Ii were attained, has no further effect

296

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

l,.

O.G

x-I

0-2

..
x

- J

600

1200 t(~)

Fig. 9.19. Time dependence of relative elastic recovery fT/1 after isothermal
extension up to the stress a = 1.3 x 10 5 Pa and elastic strain Ii = 2.6. Symbols 1 and
2 correspond to isothermal retardations at 22 and 70C, respectively; symbol 3
corresponds to the retardation under non-isothermal conditions.

on the retardation process. Figure 9.19 shows that the limiting length IT in
the non-isothermal retardation is larger than that in the isothermal cases.
This means that the non-isothermality results in a decrease in the
specimen's contraction. Note that, unlike the isothermal cases, the
principle of time-temperature superposition is not generally applicable to
non-isothermal cases.
The effect of suppressing retardation under cooling was explained in
Ref. 197 by the occurrence of tangential stresses distributed across the
specimen, whose relaxation consumes a part of the elastic energy
accumulated in the previous extension. No calculations have been
performed to confirm the explanation. When stress recovery is under
cooling conditions with the temperature below glass transition or melting
temperatures, the contraction of the specimen decreases additionally due
to the hardening effect. It is known (see e.g. Ref. 400) that extrudate swell
can be simulated by strain recovery. The results of this section show that
isothermal swelling is greater than non-isothermal swelling.

CHAPTER 10

Flows Close to Simple Shear and


Simple Extension

10.1 INTRODUCTION
Several complex problems of viscoelastic fluid mechanics, of both scientific and industrial importance, are gathered in this chapter. Special
attention is paid to easily estimating their solution and to revealing the
properties of the constitutive equations responsible for the occurrence of
new effects.
First we discuss the helical flow of viscoelastic fluids in annular dies,
which can produce an increase in output for the Poiseuille flow by the
rotation of one of the cylinders. This effect uses the shear thinning
properties of viscoelastic fluids and is discussed in Chapter 8 (Section
8.4.5) when the imposition of a pulsatile component in the pressure drop
is considered. The effect of the first normal stress differences in shearing
is very important in understanding the peculiarities of flow in a disk
extruder. In order to describe effectively the helical flow and the flow in
a disk extruder we will use some simplifications in viscoelastic constitutive equations. Both the shear and normal stresses dependences on the
shear rate are important in the design of rotary rheometers with 'optimal
geometry' as well as in the optimal design of a 'floating' thrust bearing
with variable clearance between rotating surfaces.
The chapter also establishes some easy approximate relations between
uniform and non-uniform extensions which are applied to the certification

298

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

of polymer raw materials for processing needs. The simplified equations


for the non-uniform extension are further used for the evaluation of
free and stretched extrudate swell, where the results of calculations
are compared with experimental data. Also discussed are theoretical and
experimental studies of jet withdrawal of low viscosity polymer
solutions from a jar with a free surface, with a description of hardening phenomena. As a good theoretical example, we consider the
problem of nonlinear wave propagation along a viscoelastic bar, and
numerical studies of the problem of the bar's strike against a rigid
obstacle.
At the end of the chapter the relation between simple and pure shear
flows of viscoelastic liquids is briefly discussed.
10.2 STEADY HELICAL FLOW OF VISCOELASTIC LIQUIDS
10.2.1 General Formulation
We shall study a spiral steady isothermal viscoelastic flow in an annular
die with cylinders of radii Rl and R2 (Rl <R 2 ) and length L, under the
mutual action of a positive pressure drop /1p/L and the rotation of an
internal cylinder with a constant angular speed no (Fig. 10.1).

AP

Fig. 10.1.

Sketch of helical flow.

The problem was first studied in Ref. 401 where the effect of the flow
was considered in a general form. Experimentally, the flow was first
investigated in Refs 402 and 403 for inelasic dispersed media with a
sharply pronounced anomaly of viscosity. Theoretically, the problem was
solved in Refs 404-414 by using the various rheological models, of which
the most general constitutive equation for viscoelastic liquids was a
model of a simple liquid. lol
The mutual interests in this problem were stipulated by its importance
in technical applications related to a notable increase in the output of the
basic Poiseuille flow by rotation of the cylinder. This effect has been
widely used in the extrusion of polymers and in oil drilling.

299

Flows Close to Simple Shear and Simple Extension

It is convenient to study the flow in the cylindrical coordinate system,


{r,<p,z} (z--+1, r--+2, <p--+3), directing the z-axis along the axis ofsymme-

try. From this point, the effects of the entrance and exit flows are ignored,
assuming that L~ R2 - R i . Then in the fully developed flow, there are two
components of velocity, Vz = u(r) and vq> = rQ(r), subordinated to the
non-slip conditions at the walls:
(10.1)

-,0'-0]

The strain rate and vorticity tensors are represented in the form:

e~~

u'

[:'

rQ'

o~] ,

m~~[~,+o

-u'
0

(10.2)

Hereafter the prime denotes derivatives with respect to radius r.


It is well known (see, e.g. Ref. 101) that by the orthogonal transformation with the matrix Q
J.l=rQ'/y,

v=u'/y,

(10.3)
this problem is reduced to the effective simple shear in a certain
orthogonal curvilinear coordinate system, with the effective shear rate y
and the three common viscometric functions: shear stress ,(y), and the
first 0" 1 (y) and the second 0"2 (y) normal stress differences. Then the reverse
transformation results in the following expressions for the non-trivial
components of the stress tensor (1 in the cylindrical coordinate system:
O"rz =
0"q>z

vr(y),

= J.lVO" 2 (y),

O"rq> = J.l'(Y)
O"rr - 0" zz = 0" 1 (y)

+ V 2 0" 2 (y)

(10.4)

For the particular case of a pure annular flow, J.l = 1 and v = 0, the
problem is considered in Section 8.4.2. The same formulae hold for
orthogonal steady shearing in an annular die when Ap = 0 and, say, the
external cylinder has an axial motion with a constant speed U.
The steady flow problems which can be locally reduced to simple shear
are called viscometric. 101 Helical flow is one example of such complex
flows. Another example, a possible rectilinear flow in a long tube of
arbitrary cross-section, is considered in Section 8.2.

300

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

In the following, we shall use the general formulae (10.4) which are
independent of the particular viscoelastic constitutive equation employed. All that we need here are the three basic viscometric functions,
which can be, for example, specified by the simple shear relations
established in detail for the viscoelastic constitutive equations under
study in Chapter 5.
10.2.2 Expression for the Flow Rate
Following Refs 101 and 413, this section aims to obtain an expression for
the flow rate Q as a function of the pressure drop I1p/L and angular
velocity Q o. For this purpose, as shown below, only one viscometric
function, the flow curve ,(]i), has to be known, because as in this case of
steady shearing under study, the normal stresses do not affect the flow.
This means that the results of this study are of general importance,
whether the liquid possesses shear elasticity or not.
In this section, the flow curve ,(y) is simply specified by the power law,
y= eLm, with e and m being constants. As shown below, this flow curve
describes the practically important interval of shearing very well. Then
using eqns (10.3) and (10.4) one can obtain the expressions for the shear
stresses arz and au" as follows:

rz

~
e 1/ m[(U')2 + (rQ,)2]<1-mJ/2m'

a r </>=

erQ'
1/m[(u,)2+(rQ,)2J<1-mJ/2m
(10.5)

Additionally, the inertialess equations of momentum balance are used in


the form:

op

1 0

uZ

r ur

-;- = -

-;-(l"G"rz),

assuming that all the components of the stress tensor, except the isotropic
pressure p, depend only on radius r. In eqn (10.6), 'rr is the component of
extra stress tensor. The solution of eqn (10.6) for the pressure and
tangential stresses is of the form:
p = -

Jz+h(r),

arz =

-;:- -

rJ

2'

(10.7)

where h(r) is a certain function of r, J is a given (positive) constant,


gradient of pressure, and a and b are two unknown constants, searched
for.

Flows Close to Simple Shear and Simple Extension

301

Using the power law, Y= Crrn, with


(10.8)
we can resolve eqns (10.5) relative to u' and

Q'

to obtain
(10.9)

u'=Cr rn - 1(b/r - rl/2),

Integrating eqns (10.9) with an allowance for boundary conditions (10.1)


yields:
u(r)=C

i (b/~-U/2)rrn-l(.;)d~,
r

JR1

Q(r)=aC

i rrn-l(~)d~/~3
r

JR2

(10.10)

The value of constant a is evaluated from the known value of the torque,
by the relation: M = 2nr2 a rcp L= 2naL. Additionally, I is the known
pressure gradient. Thus,
a=M/(2nL),

I=!lp/L

(10.11)

To determine the constant b in the first formula of eqn (10.10), one should
simply use the non-slip boundary condition (10.1) for the longitudinal
velocity u(r) at r = R 2 , which gives:

R2

(b/r - rl/2)rm-l(r)dr=0

R,

(10.12)

The boundary condition (10.1) for the angular velocity at r = Rl with an


allowance for the first condition in eqn (10.11) yields:
Qo =

CM lR2 rm - l(r)
- - 3 - dr
nL R,
r

-2

(10.13)

Substituting the expression (10.10) for the longitudinal velocity u(r) into
the common formula for flow rate through the cross-section perpendicular to the z-axis

Q= 2n

R2

u(r)r dr

Rl

and integrating by parts, results in:


(10.14)

302

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Thus, in assigning values of the constants a and j from eqn (10.11), one
can find the value of constant b from eqn (10.12) to obtain finally the two
gross characteristics of the flow, the dependences no(M, Ap/L) and
Q(M,Ap/L).
Using eqns (10.10)-(10.14) for m > 1, one can prove that:
(1) there is only one solution b(M,J) (f = Ap/L) of eqn (10.12);
(2) 8b/8j> 0, 8b/8M ~O;
(3) the shear stress (!rz= b/r-rj/2, considered at a fixed value of j as a
function of torque M, is limited;
(4) the longitudinal velocity u(r) always has a maximum value attained
at r=ro, where rORl +R 2)/2;
(5) at a fixed value of j, the flow rate Q and the angular velocity no are
increasing functions of the torque M.

In the case of a Newtonian liquid (m= 1), the flow rate Q does not depend
on no. When the liquid is dilatant (m< 1), i.e. possesses shear-thickening
properties, the flow rate Q decreases with growing no and the flow in the
annular gap with a rotating cylinder can be used as a locking device.
Of practical interest is the case when the inequality

Ib/r-rj /21 ~M/(2nLR2)

(10.15)

holds true. This is the case when a relatively small pressure drop can
produce a substantial flow rate at the expense of a great decrease in the
effective shear viscosity caused by the rotating cylinder. Integrating eqns
(10.12)-(10.15) under assumption (10.15) yields:
m - 1 1 - s2m - 4
2m-2'
m-s

R1
s= - ,
R2

b~jRi-2
41

nC

Q~jR14(m_3)(1-s2m-4) 2nLRi

C [ M ]m
---2 (1_s2m)
2m 2nLR l

no~ -

]m-l [

. 1-

(m-1)(m-3) 1_S2m - 4]
(m-2)2
1_s2m-2
(10.16)

Equations (10.16) show that under assumption (10.15), the flow rate is
proportional to the pressure drop and nli -11m .
10.2.3 Results of Theoretical Calculations and a Comparison with
Experimental Data
The experiments were carried out using industrial PP with the molecular
weight M ~ 3.5 X 105 at 230C. The pilot plant for the study of complex

Flows Close to Simple Shear and Simple Extension

303

shear is described elsewhere. 413 The dimensions of annular die cylinders


shown in Fig. 10.1 are: R1 =0.85, R2 = 1.0 and L= 15 cm. The cylinders
were thermostatted. The flow curve for PP at 230C displayed in Fig.
10.2 demonstrates that the power law is applicable in the region of
measurements covered by more than two orders of variation in y. The
constants of the power law found from fitting the straight line in Fig. 10.2,
are: C ~ 5.2 x 1O- 12 Pa -ms -1, m~ 3.1. Equations (10.12)-(10.14) were then
solved numerically relative to the values no and Q searched for, under
given values M and f = ApjL.

lOll-

-r(Pa)

100

Fig. 10.2.

The flow curve for PP at 230C.

Figure 10.3-10.5 demonstrate curves 1-5 relating to values of f =(0.5,


1.5, 3.0, 4.0, 4.5) X 10 5 Pa cm- l The dependences of the specific torque
MjL on the angular rotating velocity no shown in Fig. 10.3, are
monotonically increasing functions of no, which are loosely related to f
at no> 5 s - 1. Solid lines in Figs 10.4 and 10.5 represent the predictions
which demonstrate the effect of rotation on the polymer throughput Q.
The open circles in Fig. 10.4 represent the experimental data. The dashed
lines in Fig. 10.5 denote the dependences Q(n o) obtained by the approximate formulae (10.16). The comparison between the exact and approximate dependences shows that the approximate formulae produced a good

304

Nonlinear Viscoelastic Effects in Flows oj Polymer Melts

10

Fig. 10.3. Plots of torque per length unit M /1 versus angular velocity of internal
cylinder no for helicoidal flow. Curves 1, 3 and 5 denote the pressure drops
1=(0.5, 3.0, 4.5) x 10 5 Pa cm - 1.
1.0

f..f

a
3 1

cm s-

cm3 g-1
0.75

0.5

4.

8 .520 (s-1)

Fig. 10.4. Experimental (points) and


computed (lines) plots of flow rate Q
versus angular velocity of the internal
cylinder no. Curves 3, 4 and 5 are
related to the pressure drops 1=(3.0,
4.0, 4.5) x 10 5 Pa cm - 1.

o~

2.5

____ ____
~

5.0

~~~

s.!o(S-') 10

Fig. 10.5. Comparison of precise and


approximate calculations of plots of
flow rate Q versus angular velocity of
the internal cylinder no. Curves 1, 2, 3
and 5 are related to the pressure drops
1=(0.5, 1.5, 3.0, 4.5) x 10 5 Pa cm -1.

result even in the region where the values of pressure drop are fairly high.
Gapwise distributions of the effective shear viscosity ~ can be seen from
the diagrams shown in Fig. 10.6. Its value varies between the two limiting
bounds: the upper curve is close to that for the flow of polymer with fixed

Flows Close to Simple Shear and Simple Extension

305

o.05~~------'---~~---'---------'

0.85

D.lJO

t(em) 1.00

Fig. 10.6. Gapwise distributions of effective viscosity for the pressure drops
J=4.5 x 10 5 Pa cm -1. Curves 1,2,3,4 and 5 are related to the values of angular
velocity of internal cylinder no =0.87, 1.8, 3.1, 5.3, 12.9 S-l.

cylinders; the lower one to the flow between two rotating cylinders with
dp=O.
Figures 10.4 and 10.6 clearly show that the effect of a rotating cylinder is
associated with diminishing the mean effective viscosity, which in turn, results in an increase in the flow rate Q. In the study under discussion,413 it was
possible to increase the flow rate more than two-fold by the action of the
rotating cylinder at a constant pressure drop, as compared to the case
no = O. Figures 10.4 and 10.5 also show that to obtain the same increase in
the flow rate by increasing pressure drop only, requires a drastic increase in
the pressure drop, at least for the polymer considered. Therefore enhancing
the flow rate in this kind of flow by the rotation of one of the cylinders is energetically profitable. This can be also proved by straightforward calculations.
It should be also mentioned that the comparison between the calculations
and data obtained for the flow of a polymer solution with one combination
of angular velocity and average axial flow speed, is demonstrated in Ref. 410.
An increase in the flow rate in an extruding polymer through an
annular die at a constant pressure drop can be also achieved by torsional
vibrations of one of the cylinders, imposed on the longitudinal flow (see,
e.g. Refs 333,415-417).

10.3 VISCOELASTIC FLOW IN A DISK EXTRUDER


10.3.1 Scheme of Disk Extrusion
Consider the steady flow of elastic liquid (1) in the gap of width h between
the two coaxial disks of radius R1 (Fig. to.7). The upper disk (2) rotates

306

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Fig. 10.7. A sketch of a disk extruder. 1, Steady flow of elastic liquid; 2, rotating
upper disk; 3, immobile lower disk.

with a constant angular velocity n around the z-axis, the lower disk (3) is
immobile. There is a hole with radius R2 in the central part of the disks.
Under the action of normal stresses arising in the main shearing flow in
the gap and directed to the centre, a secondary centripetal flow occurs
which results in the extrusion of the liquid through the central hole with a
flow rate Q. The secondary flow of an elastic liquid is steady if the liquid
is fed from the periphery of the disks. The name 'disk extrusion' came
from the polymer processing industry where the machine is sometimes
used for pumping or feeding polymer melts. It is remarkable that at
similar hydrodynamic conditions, the flow of a Newtonian liquid is
centrifugal under the action of inertia forces.
In these machines, the inequalities between dimensions: Rl ~ R2 ~ h
are commonly satisfied. Thus as the first approximation, the flow of
elastic liquids can be considered as that between the disks of unlimited
radius.
Below we consider steady flow within the inertialess approach. The
main objective of the theoretical treatment is to analyse the secondary
centripetal flow at relatively small angular rotation speeds n, when
non-isothermal effects are negligible. At the end of the chapter we discuss
the phenomena occurring in disk extrusion at arbitrary values of n.
10.3.2 Formulation of the Problem
We proceed from the single Maxwell-like model discussed in Chapter 1
(see eqns (1.9a), (1.42) and (1.51)):
c-c. Vv -(Vv)t. c+ [c 2 -C(ll - 12 )/3 -0]/28 = 0
detc= 1,

11 = trc,

12 = trc- \

t1

= - po + (p/n)c n (10.17)

where c is the elastic strain tensor, t1 is the stress tensor, 0 is the unit
tensor, v is the velocity vector, and /1, {} and n are the elastic shear
modulus, relaxation time and the power factor, respectively, which are

Flows Close to Simple Shear and Simple Extension

307

considered below as material constants. In the following, we will use eqns


(10.17) for the case of steady flow when dimensionless angular speed
08~ 1. This case is examined in Ref. 418 by using UCM model. The
following calculations were performed by A.I. Semdyanov.
It is convenient to consider the problem in the cylindrical coordinate
system {r, cp, z), where r+-+1, cp+-+2 and z+-+3. Let us represent the tensor
c in the form: c = 0 + 8. It is easy to note that for the flow under study, the
components e",z, 8 r ", and v", are odd while the other components of 8 and v
are even functions of parameter 08. Then within the order 0(0 2 8 2 ) we
find:
detc ::::; 1 + (11 - 12 )/2,

(Il - 12 )/2= 8rr + 8",,,, + 8zz-8;",-8~z =0


(10.18)

Now, disregarding the terms of order higher than 0(Q 2 82 ) and using eqn
(10.18) and flow symmetry over 08, we can write eqns (10.17) in the
component-wise form:
20 r vr - (28 rr + 8;",)/(28) = 0,
Or Vz + OzV r - (28 rz + 8r",8 rz )/(28) = 0

(10.19)

vr/r + OzV r + (Or V'" - v",/r)er", - (28",,,, + 8;", + 8~z)/(48) = 0


OZV", -8",z/8= 0,
(1

= -

OzV z -(28zz +8~z)j(48) =0

pO + 118

(10.20)

Continuity and inertialess momentum balance equations are:


OzV z + r-1or(rvr) =0,
(10.21)

Equation (10.17)-(10.21) represent a closed set, with the following natural


boundary conditions:
z=O: vr=v",=vz=O,
p(Rd=p(R2)::::;0

(10.22)

10.3.3 Solution of the Problem


The solution of eqns (10.19) under boundary conditions (10.22) has the
form:

Vz=O

308

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

8rr

= _

2~b (~: _~).

(10.23)

1 o. 202r2

o.rO

8q>z=h

8zz=-2~'

where b is a constant of integration. The components of the velocity


vector also satisfy the first continuity equation in the set (10.21). Substituting the components 8ij into the momentum balance equations (10.21)
yields:
]
2bO
r
30 20. 2
p=p(r)=J.1 [ V ln R2 - 4h"2 (r2-Rn

(10.24)

where the condition p(R 2 )=0 has been taken into account. Using the
second condition p(R1)=0, gives the value of the constant b:
b=

~o.20 Ri-R~
8

In(RdR2)

(10.25)

The radial flow rate for disk extrusion is defined as follows:


Q= -2n Soh vrrdz=(nj8)Oo. 2h(Ri-RWln(RdR2)

(10.26)

It should be noted that if there is a pipe flow above the hole in the disks,
the pressure at the hole does not vanish, i.e. p(Rd = Po # O. Here Po should
be determined by equalizing the flow rates in the pipe flow and the value
of flow rate in the disk extruder. In this case, the value of constant b will
be changed.
10.3.4 High Intensity Disk Extrusion

Solution (10.23) can be considered as the sum of the main shear flow
without a hole (b = 0) and a disturbance produced by a hole and
represented as a secondary radial flow. Then the terms containing
positive powers of radius r will correspond to the non-distributed flow
between disks, while the terms with negative powers will correspond to
the disturbance due to the presence of the hole. This representation seems
physically quite natural, since a small hole in the centre of disks will
hardly cause a substantial flow at a large distance.
When o.O~ 1, the problem of disk extrusion, using a single UCM
model as an example, was considered in Ref. 418. It was shown there that

Flows Close to Simple Shear and Simple Extension

309

the disturbance of viscometric flow between disks due to the hole can be
presented as the sum of series over the inverse powers of radius. As a
result, the flow rate was found to be in a good agreement with the
experiments,419 as a limited function of the rotation speed and geometrical parameters. The results of the comparison between calculations 418
and experiments419 are presented in Fig. 10.8. In principle, there may also
exist other causes restricting the flow rate in disk extrusion: dissipative
heating and melt flow instabilities.

'I
3
2

6.-1

0-2
.-3

Fig. 10.S. Theoretical (lines) and experimental (points) plots of flow rate
versus the number of revolutions fi in disk extrusion (R1 = 3.5 cm, R2 =0.07 em).
Curves 1, 2 and 3 are related to the gap thicknesses h = 3.2, 1.6 and 0.8 mm,
respectively.419

Since in steady disk extrusion, the radial flow rate Q defined in eqn
(10.26) is a constant, all the components in radial velocity Vr with the
powers of radius other than j = -1 do not contribute to the flow rate.

Therefore these components in radial velocity Vr are related to circulating


flows, which are important for mixing. This is why disk extruders have
found applications in industry as mixers and dosators.
It should be noted that the radial flow with a constant flow rate also
occurs in the disk extruder when the rotating disk performs rotating
oscillations.
The problems of disk extruder design and operations are discussed
elsewhere. 334 ,419-423

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

310

10.4 ON OPTIMAL GEOMETRY FOR GENERALIZED


ROTATIONAL FLOW OF VISCOELASTIC LIQUIDS
IN THIN GAPS
This section studies, following Ref. 424, some problems that are related to
the optimal design of the geometry in rotary rheometers and thrust
bearings, which is based on such optimum criteria as minimum consuming power and maximum thrust.
10.4.1 Basic Relations and Formulations
Figure 10.9 demonstrates a rotational shearing flow of a viscoelastic
liquid in a set with a thin gap whose median surface is a surface of
revolution. Here zo(r) is the equation of the median surface, ho(r) is the
gap's thickness, and r is the distance of a point M disposed on the median
surface from the axis of revolution. It is well known that both principal
curvature centres, 0 1 and O2 , of a surface of revolution are disposed on
the same normal to the surface in the point M, intersecting the axis with
the angle ({J; the second curvature centre O2 being situated on the
revolution axis. The principal radii of curvature, Rl = I0 1 M I and R2 =
IO2 M I, as well as angle ({J in Fig. 10.9 are defined by the familiar
relations:
R 1(r)=r)1 +z'rl/z'o, cos ({J= 1/)1 +Z'o2

(10.27)
In addition, there is a simple relation that may be easily derived from
eqns (10.27):
Rl cos ({J d({Jldr= 1

(10.28)

z
.....- - r , - - - - - > - !

1-_---:.~I:..------1---,l--r

Fig. 10.9. Geometry of a thin gap on


a surface of revolution. 424

311

Flows Close to Simple Shear and Simple Extension

In the following, the generator of the median surface will be designated as


the meridian, the circle of radius r passing through the median surface will
be the parallel, and the gap's thickness will be counted off along the normal
to the meridian. In each point of the smooth median surface, it is feasible to
introduce a local orthogonal curvilinear 'visco metric' coordinate system
Xl> X2, X3, where Xl is directed along the tangent to the parallel.
Now, let us consider a steady inertialess shearing flow of a viscoelastic
liquid produced, for example, by a revolution of an internal surface that
rotates with constant angular velocity w relative to an immovable
external surface. In accordance to the main assumption, the flow of the
liquid will be treated in the frames of lubrication approximation. This is
where the fields of shear rate and stresses are assumed to be locally
homogeneous, the variations of ho(r) are assumed to be smooth, and the
thin shell of revolution formed by the gap is assumed to be 'gently
sloping'. The latter means:
(10.29)

In the local coordinate system, the velocity vector


represented as:

V={Y X 2,0,0},

In

the gap is

Y= wr/ho(r)

(10.30)

where y is the local shear rate, and the local stress system corresponding
to the simple shear is:
N 2(y) = 0"22-0"33

(10.31)

where 0"12 is the shear stress, O"ii are the normal components of the stress
tensor, and Nl(y) and N 2 (y) are the first and second normal stress
differences.
Along the meridianal direction X3, the equilibrium condition for a
small rotating liquid element is:
d/dcp(0"33 hoR2 sin cp)- 0" 11 hoRl cos cp - 0"22R2 sin cp dho/dcp =

By using the relation R2 sin cp = r along with eqns (10.28) and (10.32), this
equation can be rewritten in the form:
(10.32)

Another equilibrium condition of the liquid's element along the direction


normal to the surface of revolution with provision for the 'gently sloping'

312

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

condition (10.29) brings us to the conclusion of local homogeneity of the


stress field in the gap. This usually takes place in the flows under
consideration within the frames of lubrication approximation. After all,
the condition of torque equality for a thin ring gives:
dM = 2nr2(J 12 ds,

ds

= J1 +z{?

(10.33)

where dM is the elementary torque for a ring with generator's length ds


along the meridian.
From eqn (10.32), it is easy to obtain the distribution of the gap's
normal stresses along the radius (or meridian):
(Jll = - Po + N 1(-Y) + N 2('Y)-

f2

(J22= -Po+N 2(y)-

+N2+xN2h~(x)/ho(x)J dx/x

f2

[N1

[N 1 + N 2 +xN

2h~(x)/ho(x)J dx/x

(10.34)
where Po is a certain external pressure given at r=r2'
The total thrust produced by the liquid flowing through the gap is
T= -2n

r2

f(J22 cos qJ ds

r1

After substituting here the expression for (J22 from eqn (10.34) and using
the relation cos qJ. dqJ/dr = 1 resulting from eqns (10.27), the latter equation can be represented in the form:
T=n(d-ri)po+n

f ro [(r-ri/r)(N1-Ndr/Y}-2N2/rJ dr
r1

(10.35)

The formulae for the torque M, the gap's volume V, and the area of
median surface F are of the form:
(10.36)

F=2n

f'1 rJ1 +z{? dr


r2

(10.37)

Flows Close to Simple Shear and Simple Extension

313

Additionally, it is easy to prove that for the steady rotary shearing of


viscoelastic liquids, the torque M is connected with total dissipation D by
the evident relation
D=MOJ

(10.38)

Formulae (10.35)-(10.37) coincide with those obtained in Ref. 425 in


the case when r1 =0.
From now on we will treat the quantities T, M, V and F as functionals of two functions ho(r) and zo(r), where T and M also depend
parametrically on the given angular velocity OJ. This means, in particular, that at the fixed volume of the gap V, angular velocity OJ, and
the liquid's rheology, we can vary the thrust T and torque M only
by variations of geometrical distributions zo(r) and ho(r). Thus,
some typical variational problems arise, such as how to design
better the rotary devices bearing in mind some goals (or the goal
functions).
Two possible and practically valuable variational problems were under
study in Ref. 424:
(i) to find the geometry, ho(r) and zo(r), of a rotary device which gives
a minimum torque M under fixed volume V; and
(ii) to find the geometry which gives a maximum thrust T under fixed
volume V.
It is natural to initially attempt treat the problems as standard
isoperimetric problems of calculus of variations. Below we discuss the
main results of the analyses. 424

10.4.2 Isoperimetric Problems of Extremum of Torque

A tedious but standard analysis of the first problem (i), where both values
r1 and r2 were also varied, resulted in very general and distinct conclusions.
For a certain elastic liquid, there are only two local solutions of the
variational problem: minzo,hoMI V=const.:
(a) The regular solution
zo(r)=O;

ho(r) = rep

(10.39)

corresponding to a biconical plate rotameter with certain given


values of geometrical parameters, small angular gap ep and radii r1

314

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

and r2. The following familiar formulae are valid:


M=O"dy)V;'cp;

V=(t)n(r~-ri)cp;

y=w/cp

T= n(r~ -ri)[po + N 1(Y)!2] + nrHN 1(y)-2N 2(Y)]ln(rdr2)

(10.40)

(b)The peculiar solution


zo(r) =

ctJ(O<z~L),

ho(r) = constant

(10.41)

corresponding to a standard Couette cylindrical rotameter with


certain given values of geometrical parameters ho, Land r1 =r2.
The following well-known formulae are valid in this case:

(10.42)
It is seen that the Couette rotameter delivers an absolute minimum for
thrust T which is T = O. Also, at the same values of shear rate y and
volume V, the value of torque M provided by both the late biconical and
cylindrical Couette rotameters, are the same and minimal between other
possible configurations of the gap.
In practical rheometry, the Couette cylindrical rheometer is usually
provided with a bottom. An example of the rotary rheometer of combined geometry is shown in Fig. 10.10. If the combined rheometer is
designed to consume a minimum energy (torque), the previous consideration shows that 10 = 0, the parameters Land R, limiting the size of the
instrument, should be given, and the question remains about the relation
between the values of gaps, cp and h/R. If we consider further the problem
of minimum torque for the combined rheometer

~~x

Fig. 10.10. Sketch of a cylinder-cylinder rheometer with a bottom of bi-conical geometry.424

315

Flows Close to Simple Shear and Simple Extension

under a given (fixed) value of the total volume of the gap,


V = 2nR(hL+ RZq;/3) = constant

the problem will be reduced to the standard problem of the conditional


extremum for M considered as a function of two variables, q; and h/R,
which solution is: q; = h/R. This is a continuous conjugation of the two
configurations.
The examples of calculations 4z4 show that the difference in values of
torque M for the rotameters of various geometries is relatively high at
small shear rates and decreases with a growing shear rate. Other
variational problems of the rational design of a rotary viscometer
demanding minimum torque can also be formulated. 4z4
10.4.3 On Maximization of Thrust
The isoperimetric variational problems (10.35)-(10.37) on the extremum
of thrust T under a given volume V of rotameter were also extensively
studied in Ref. 424. The analysis showed that a local extremum of the
thrust functional is related to a local minimum for the solution of
variational problems, depending on the angular velocity OJ and rheological properties of viscoelastic liquids. As mentioned previously, the absolute minimum of the thrust, T = 0, is achieved in the Couette rheometer
whose geometry corresponds to a peculiar solution of the variational
problem. Moreover, it was proved that depending on the geometry,
rheological properties of liquid and angular velocity OJ, the functional T
can take arbitrary large values, i.e. it is unbounded from above.
In order to study these effects we need to discuss in more detail the
properties of three material functions (10.31) in the simple shear. First of
all, these functions can be represented in the dimensionless form:
0"12

= Ge,,(y),

N1=GeO"l(Y)'

Nz=-GeO"z(y)

(y=yO)

(10.43)

where Ge and 0 are the high elastic modulus and the mean relaxation time
defined in Section 1.9, and" and O"j (O"j>O) are dimensionless shear and
normal stress differences. In order to avoid using special viscoelastic
models we assume the following general properties of the viscometric
functions a small and large values of dimensionless shear rate y:
O"j(Y):::::CjY Z
O"l(y):::::ay

(0<f3~r:x<

(CZ/Cl <1);
O"Z(y):::::aly

a;

1,

a1 ;

0<r:x 1 ~r:x,

,,(y):::::y
,,(y):::::byfJ

O<al ~a)

(y~l)

(1O.44a)

(y~

1) (1O.44b)

316

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

We assume that the inequalities:


[yl+na;]'>O
(10.45)

Vn=const>O, yE [0, (0):

also hold true. The two power asymptotes (10.44) at low and high values
of shear rate, as well as the ineqns (10.45), are usually well fulfilled in
reality. As an example, Fig. lO.ll(a) and (b) display empirical plots 426 for
N 1 and a 12 versus generalized shear rate 110 y that are represented in
double logarithmic coordinates. The data of Fig. lO.11(a) for PIB P-20
and those of Fig. 10.11(b) for a butyl rubber distinctly show that in both
cases, there is an overlapping range of applicability of the power
asymptotes (10.44) for N 1 and a12, which reaches almost to two orders of
shear rate. According to Fig. lO.11, for PIB, a ~ 0.60 and p~ 0.30, and for
butyl rubber, a~0.69 and p~0.25. Unfortunately, the absence of data 426
on the second normal stress difference and also, on the constants 110 and
Ge did not allow for the determination of the values a, b, Cl and C2 in
asymptotic presentations (lO.14).

5
6
7
8
Ig (l1o~ )[dn/cM']

Ig (l1o~ )[dn/cM']

Fig. 10.11. Plots of the first normal stress difference N d2 and of the tangential
stress 0"12 versus the generalized shear rate 'loY according to Ref. 426: (a) PIB P-20
(C(~O.60, p~O.30); (b) butyl rubber (C(~O.69, p~O.25).424

Now we consider a practical problem that is related to the optimum


design of such rotational sets as disk extruders, thrust bearings, etc.
(when the internal boundary rl =0), to determine a gap's geometry in
which as great a thrust as possible could arise under minimum torque
and given gap volume. Though the results 424 show that such a problem
has no exact solution, we still can find an approximate '8-solution'
for which the torque exceeds the minimum value with the given
accuracy of 8, but the thrust is theoretically infinite. It is evident that

317

Flows Close to Simple Shear and Simple Extension

one should search for such a configuration with the plate geometry
(zo = 0) where torque is minimal, and in the region where the shear rates
are high enough. To simplify the results we use the asymptotic formulae
(10.44b) for the basic visco metric functions (10.43). In this case, one
can represent the basic functionals (10.35)-(10.37) in the asymptotic
form:
T-ndpo
2G
nr2 e

V
nr2

a(ill e/J v )a.J y,

J V =-23

(10.46)
where dimensionless functionals J y and J m depend only on the geometry of the gap and have the form:
J y=

(K(X) = x/(h(x)J J

X[K(X)Y dx,

(10.47)
where x=r/r2 and h=ho/r2 and the volume constraint is now written
as:

X2 dX/K(x) = 1

(10.48)

From eqns (10.47) one can easily perceive that the only method to
attain the increase in J y without essential variation of J m is making a
distribution k(x) singular a x = O. This distribution should be localized
in a narrow vicinity of the axis of revolution (0, xo) and the rest of the
interval's part (xo, 1) should utilize the plate biconical geometry for the
gap, which corresponds to the minimum of J m . In this case, we additionally demand:
(10.49)
where the value 1/3 1 + P corresponds to minJm with Xo = 0 and
arbitrarily small positive number.
In so doing, we can accept the singular distributions:
K*(x)=v'f(x),
f(x) =

{i

xo / x )2/a

is an

h*(x) = X' f(x)/v


(O<x~xo)
(xo~x~l)

(10.50)

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

318

where the quantities v and Xo are determined from eqn (10.48) and in eqn.
(10.49), and are given by
v(

Xo

)= 3ct+2(1-x6)
3(3ct + 2) ,

J m == vP [t + ctX6/(3ct - 2f3)]

<(1 + 8)/3 1 + p
(10.51)

From relations (10.51) one can determine the values of parameters v and
Xo depending on values of ct, f3 and 8. For the cases of PIB P-20 and butyl
rubber, discussed above, the results of these and following calculations
are represented in Table 10.1.
Table 10.1 Numerical values of parameters in eqns (10.49)--{lO.52) for rheological
parameters (;( and f3 (according to Fig. 10.11)
Parameters

Xo

J1' = -A ln2+B
A

Jm

min J m

(;(=0.60

f3=0.30

0.05

0.536

0.306

0.0326

0.398

0.252

0.240

0.05

0.657

0.287

0.0468

0.295

0.266

0.253

(;(=0.69

f3=0.25

If distribution (10.50) for K*(X) is substituted into functional J 1' defined


in eqn (10.47), it produces a theoretically infinite thrust. This result was
interpreted in Ref. 424 as the appearance of a small clearance at x = 0,
with the dimensionless value 11 = b/r2. This clearance between two rotating surfaces is due to the action of normal stresses arising in sheared
elastic liquid. As a result, the thrust will take a certain finite value in
accordance with the balance of axial forces. In a simple case, such a
process takes place if one of the surfaces is freely pressed to another. If,
however, these are fastened, then the clearance will be formed as a result
of elastic deformations in the set's construction. Thus, in the general case,
the set with the profile considered will be adopted to take arbitrary loads
due to the formation of the 'floating' (self-controlled) clearance. The
analysis of the rotating flow of elastic liquid in a freely-floating thrust
bearing resulted in the simple expression for J 1' :
(10.52)

Flows Close to Simple Shear and Simple Extension

319

where constants A and B are expressed through the parameters a, v and


Xo. Their numerical values are presented in Table 10.1 for both the PIB
P-20 and butyl rubber.
It should be noted that this hypothetical floating thrust bearing was
never tested. Additionally, at high shear rates, melt flow instabilities can
appear which reduce the effect of high thrust.

10.5 WEAKLY NON-UNIFORM EXTENSION OF


VISCOELASTIC LIQUIDS
This section analyses weakly non-uniform uniaxial elongation flows of
viscoelastic liquids, which are almost uniform in cross-section and inhomogeneous along the longitudinal direction. 427 Though for the sake of
simplicity, only the extension of viscoelastic jets of circular cross-section
are considered, the results of analyses can be easily extended to the jets of
other cross-sectional configurations.
Figure 10.12 demonstrates a scheme of experiment for non-uniform
steady extension. A cylindrical viscoelastic sample (2), adherent to the
roller (1) of radius Rb rotating with angular speed Q1, is detached from
the roller and stretched under the action of a constant force F. While
stretched, the sample is floating. In the particular case shown in Fig.
10.12, the force was assigned by a constant load with a thread passed over
the roller, as in experiments with uniform extension (see Fig. 4.10).
Technically, non-uniform extension can be carried out more easily with
liquid extruded from capillary and stretched by a force F (see Fig. 4.17).
1

Fig. 10.12. Sketch of device for non-uniform extension under constant stretching force F: 1, large roller; 2, polymeric sample; 3, bath with thermostatting
liquid; 4, mobile clamp; 5, float; 6, thread; 7, small roller; 8, load.

320

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

But in this case, shear stresses in the capillary exit may distort the
extensional flow much more than that shown in the scheme in Fig. 10.12.
It is convenient to consider the extension in the cylindrical coordinate
system with z-axis directed along the elongated sample and originated at
the point z = 0 where the sample is detached from the roller. We denote
by R(z) and v(z) the radius and longitudinal velocity of a polymeric jet
extended by the stretching force.
Two main assumptions are usually made when considering weakly
non-uniform stretching (see e.g. Refs 294 and 428):
(i) the total time derivative d/dt in viscoelastic constitutive equations
is approximated by the expression
d/dt ~ v(z) d/dz

(IR'I=ldR/dzl ~ 1)

(10.53)

(ii) with the accuracy of OCR'), the kinematics and stress distribution
are almost the same as in the case of homogeneous simple
elongation.
Although the effect of surface tension is usually disregarded, it can be
easily incorporated into analysis when working with a small viscosity and
extremely thin jets.
The above two assumptions mean that in this approach, all the tensors
in the viscoelastic constitutive equations, the stress (1 and the strain rate e
tensors in particular, can be considered as diagonal. Then the total
longitudinal components of these tensors are defined as
(F = const),

K(z)=dv/dz

(10.54)

and the velocity components are represented as follows:

J:

vz=V(Z)~VI + K(~)d~,

vr ~ - K(z)r/2,

(10.55)

where VI = v(O) is the initial value of the longitudinal velocity. It is also


assumed that in the experiments sketched in Fig. 10.12, the polymer
sample on the rotating roller is non-deformed. The correctness of this
assumption is discussed later.
Now let us establish an important relationship between uniform and
weakly non-uniform extensions. In accordance with the above two
assumptions of the non-uniform extension, each cross-section of a
stretched sample moves along the z-axis as in uniform extension under
the force F, having a time t = 0 (z = 0) the radius Ro. Then the correspond-

Flows Close to Simple Shear and Simple Extension

321

ence between the stretching time t in uniform extension and the position z
of a cross-section in non-uniform extension is given by:
z=

J;

Vet)

d, =

V1

J;

a(,) d,

(10.56)

where a=(Ro/R)2 is the relative elongation of the sample (see eqn (4.38)).
This expression is obtained from eqn (10.53) with allowance for
(10.57)
where Q is the flow rate and for the particular scheme of extension shown
in Fig. 10.9, the speed V1 of polymeric jet at z=O, V1 ~Q1R1. As follows
from eqns (4.37) or (10.57), the profile R(z) is related to the relative
deformation a(t) as
R(z) = Ro [a(t)] -1/2

(10.58)

To check eqn (10.58) one should use the dependences a(t) and R(z)
obtained from corresponding uniform and non-uniform extensions under
the same given stretching force F.
The residual length Zr in non-uniform extension is now defined as the
length that is achieved after total contraction in the elastic recovery
process (0- = 0) for the stretched section of the sample, which is simultaneously cut at the cross-sections with the positions 0 and z. The relation
between the appropriate values of Zr in uniform and non-uniform
extensions is as follows
(10.59)
where Ii is the elastic strain in uniform extension under a given constant
force (see eqn (4.40)). In deriving eqn (10.59), the formulae (4.40), (10.53)
and (10.57) were utilized. In order to calculate the dependence Zr(z) one
should use the experimental data on uniform extension under F = constant (see, e.g. Figs 6.6, 6.7 and 7.5) along with eqns (10.56) and (10.59). As
seen from these relations, Zr#z/Ii(z); the latter expression being even
independent of V1.
To compare the non-uniform steady and uniform non-steady extensions under a given constant value of stretching force F, a special
experiment with non-uniform steady extension was carried out using the
same PIB P-20 that was employed for the uniform extension (see Chapter
6). At 22C, its Newtonian viscosity '1 and shear high elastic modulus Ge
have the values: '1 = 1.1 x 106 Pa s -1 and Ge = 1.57 x 10 3 Pa.

322

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The experiment with uniform extension whose results are used for
calculations, was carried out with values (Jo=F/S o =6.3x10 3 Pa (So is
initial cross-sectional area of the sample) on the device shown in Fig. 4.10
and was discussed in detail in Section 6.3. The stress (J 0 was assigned by
the load F = 0.18 N in weight. The initial length of the sample was varied:
lo = 2-5 cm; the initial radius being Ro ::::;0.3 cm. The plots 8= l(t)/lo and
&(t) = l(t)/lr(t) which did not depend on lo are shown in Figs 6.6 and 6.7 by
open triangles.
In experiments with non-uniform extension, carried out with the device
shown in Fig. 10.12 with a roller ofradius R1 =4 cm, the values Ro and F
were found to be the same as in the uniform extension; initial velocities V1
of the sample being varied as: v1 = 2.13 x 10 - 4, 1.07 x 10 - 4 and
0.54 x 10 - 4 m s. -1 The polymeric jet's profiles were registered by photography on a 1: 1 scale. During measurements of parameter Z" the time of
elastic recovery was found to be the same as in homogeneous extension.
The extension under study was steady. But it was found that the
polymer on the roller started to deform slightly at a distance of 0.5-1 cm
before the point z = O. This pre-deformation slightly increased with a
decrease in value V1.
The curves 1 and 2 in Fig. 10.13 demonstrate the plots R/Ro and Zr/V1
versus Z/V1, respectively, re-calculated from the experimental data on
uniform extension shown in Figs 6.6 and 6.7 by using eqns (10.56), (10.57)
and (10.59). The various points relating to different values of V1 show the
experimental data on non-uniform extension. It is seen that the experimental points form the single curves which do not depend on values V1' A
good agreement between the two sets of data is observed in the region
z > 0, with a certan discrepancy in the vicinity z = 0, seemingly attributed
to the pre-deformation in the region z < O. The derivative of jet's profile,
dR/dz= -RoK/(V182 / 3 ), reaches the value -0.05 which is higher than
those commonly found in practice.
Besides, the dependence t(z) according to eqn (10.56) was also checked
directly. During the experiments, a mark perpendicular to the z-axis
was made on the sample, using paint, and then the motion of the mark
was observed in time along the axis using the initial condition tl z = 0 = O.
Open symbols in Fig. 10.13 denote the data t as a function of Z/V1 for
various values of V1. The maximum values of z measured in experiments
with above values of velocities V1 were Z= 12.7, 6.35 and 3.2 cm. The
radius of the sample, R, decreased more than three-fold on these
lengths. Curve 3 in Fig. 10.13 shows the dependence t(z/vd re-calculated
from Fig. 6.6 using eqn (10.56). It is seen that the function t(z/vd

323

Flows Close to Simple Shear and Simple Extension

R/Ro

0-. -1

'I 0,5

ll-.-2
0-. -3

If

0
5" Z IV; fO-zCd')

Fig. 10.13. Non-uniform extension of PIB melt at constant force with


= 6.3 X 10 3 Pa. Computed (lines) and experimental (points) plots of relative
radius RjRo of specimen (1), its residual length Zr related to the initial velocity
VI (2), and the time t of motion of a fixed sample's cross-section versus
longitudinal coordinate z related to the initial velocity VI (3) at various values of
VI' Curves 1, 2 and 3 are related to VI = (2.13, 1.07, 0.54) x 10- 4 m s -1.
(To

does not depend on VI' With the minimum value vl=0.54x 1O- 4 ms-1,
a small deviation between the points and the curve is observed, which
is presumably due to the pre-deformation of the sample on the
roller.
Now let us consider another case when non-uniform steady extension
is achieved not by the action of a constant force but by the second roller
with radius R2 = Rj, rotating with angular velocity O 2 (0 2 ) Od; the
distance between the roller's centres being equal to L (see, e.g. Ref. 429).
The scheme of the experiment is shown in Fig. 4.11(a). It is commonly
accepted that in such an instrument, the 'elongation viscosity' defined as
rye) = a(L)/K(L) coincides with that defined in the steady uniform elongation flow. Using eqns (10.28) and (10.31) one can quickly establish,
however, that in this steady non-uniform extension, the following formulae
K=dv/dz

(10.60)

324

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

hold true. Therefore (J(L) = (J OV2/Vb K(L) = dv/dz IL and expressions (10.60)
show that the value of K(L) is not steady. This means that deriving
elongational viscosity from such an experiment is not generally correct.
Formulae (10.53)-(10.60) which relate uniform and non-uniform extensions allow us to analyse non-uniform extension which usually prevails in
polymer processing, by means of well defined rheological methods
developed for the uniform extension of polymer melts. In particular, these
methods allowed us to show the following: 40o
(i) drawing of a polymer from a capillary with allowance for deformation in polymer extrudate can be simulated by a uniform
retardation with additional extension under F = constant superposed on it (see Section 10.7);
(ii) the extrudate swell in non-isothermal conditions, with the temperature varying inside and outside capillary, can be substantially
decreased as compared to the isothermal case.
Finally, let us briefly discuss the possible manifestation of the hardening effects in non-uniform extension ('flow slow-down'), which are studied
in Chapter 7 for homogeneous extension. It should be remembered that
this effect was observed in the region of highly nonlinear deformation for
LDPE and was not observed in the simple elongational flow of PIB. The
plots of R/Ro versus Z/Vl recalculated from known dependence i(t) (Fig.
7.6) by using eqns (10.56) and (10.58), are presented in Fig. 10.14. Curve 1

1.0

0.8

'-

o ~------~------~--------~------~~r---~--o

10

13

20

-2

40

z/V; X/O (;I)

Fig. 10.14. Plots of dimensionless profiles RjRo versus ratio of longitudinal


coordinate z to the initial velocity Vl of sample, recalculated from the uniform
extension data at a constant force with (Yo = 104 Pa: curve 1, LDPE; curve 2, PIB.

Flows Close to Simple Shear and Simple Extension

325

in the figure denotes the jet's profile R(z) for polyethylene at 125C; curve
2, for polyisobutylene at 44C. As noted before, at the indicated temperatures, the two polymers show almost the same rheological behaviour in
the linear region of deformation.
The profiles shown in Fig. 10.14 correspond to a polymer extrudate
exiting from the capillary with the mean flow velocity Vl and extended by
a constant force F; the extrudate swell being neglected. Because of
the hardening phenomena, the elongational rate K(t) for LDPE is
much lower as compared with that of PIB, with the same values of Vl
and F. Figure 10.14 demonstrates that this results in the fact that the
same value of radius is attained for LDPE considerably later than with
PIB.

10.6 CERTIFICATION OF RAW MATERIAL FOR THE


POLYMER PROCESSING INDUSTRY
In the plastic processing industry, shear tests such as the 'index of melt
fluidity' are commonly used for rapid evaluation of polymeric raw
materials. In this test, the mass of polymer is extruded from a capillary
rheometer of a standard geometry under the action of a standard load
during a standard time, say 10 min; after which some very important
conclusions about the technological processibility of the material are
usually drawn.
However, as have been shown in Chapter 7, there are some polymers,
e.g. low-density polyisobutylene and poyethylene, which demonstrate
almost the same rheological behaviour in simple shear but are still
appreciably different under simple extension. This means that there is a
need for a quick and robust test to evaluate the rheology of rough
polymeric materals under simple elongation.
One of the devices for processing needs proposed by Meissner was
further elaborated in Ref. 430. It consists of rollers drawing a polymer
extrudate out of a nozzle with a greater speed than the speed of extrusion;
the stretching force being measured in time. This drawing is usually
carried out non-isothermally, under natural conditions of extrudate
cooling. This non-isothermality, along with variations in strain and
thermal histories released to the deformation of polymer in the precapillary region, makes it very difficult to interpret the results. Seemingly
for this reason, it is also difficult to compare the data obtained in different
devices. 43o

326

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

As an additional elongational test, Refs 431 and 432 propose the use of
uniaxial uniform extension under constant force. As has been discussed in
Section 4.5, this test is simply to carry out and keep under isothermal
conditions. Additionally, the results of the test are easy to interpret and
re-calculate, by using formulae (10.53)-(10.60), to the non-uniform extension, and this test is the closest to real polymer processing. In this test, the
time to attain a fixed relative elongation is measured in the extension
under a standard load.
The calculations and experiments discussed in Chapters 6 and 7
demonstrate that unlike the simple extension with assigned kinematics,
the extension under a given force is very sensitive to the dependence of
stress on the recoverable deformaton, i.e. to the form of elastic potential
(see, e.g. Fig. 6.16). This also speaks in favour of applying this regime of
extension to the certification of raw polymer materials.
Also, this method proved to be sensitive to changes in the molecular
mass, molecular-weight distribution and branching ratio of macromolecules. Using uniform extension at a constant force it is also possible
to predict the discontinuities on a polymer extrudate (see Chapter 11)
which may occur in forming fibres and films. Numerous examples of the
certification of polymeric raw materials on the basis of extension under a
constant force are given in Ref. 432.

10.7 ON THE STRETCHING AND SWELLING OF AN ELASTIC


LIQUID EXTRUDED FROM A CAPILLARY DIE 433
It is well known that a polymer extrudate leaving a circular die, swells in
the diameter due to the release of elastic energy accumulated in capillary
flow. This phenomenon, which is of high engineering importance, has
been studied in numerous works and is variously referred to as swelling,
die swell or the Barrus effect. To the authors' knowledge, the latest
detailed measurements of the effect can be found in Ref. 434. However, in
practice, especially in the polymer fibre-spinning process, the swelling
takes place whilst the polymer fibre leaving the die is being stretched
along its axis. Such a process has been studied for isothermal conditions,
e.g. in Ref. 435, as well as in numerous industrial research papers. The
main feature of the tensile experiment is that after swelling just beyond
the die outlet, the polymer extrudate becomes thinner downstream due to
the action of the stretching force. The greater the stretching force, the
smaller the amount of swelling, and if the stretching force reaches a

Flows Close to Simple Shear and Simple Extension

327

sufficiently large value, the polymer may even be pulled off the capillary
wall. 436
A theoretical description of the problem without any simplifying
assumptions is extremely difficult to obtain, even numerically, and so
several approaches have been taken in the literature to enable us to
obtain at least some simplified numerical results. A simplified approach
for calculating the swelling ratio without stretching was proposed by
Tanner (see, e.g. Ref. 4). The theoretical description of the swelling ratio
under stretching was proposed in Refs 437 and 438, taking into account
the polymer flow both inside and outside the capillary. The solutions
were matched by equating the elastic energy flux at the die outlet.
However, their results can only be treated as semi-empirical because
outside the die, the extrudate was considered to be purely elastic (thus not
permitting the computations of the jet's profile) and in the internal
capillary region empirical data were employed to calculate the values of
the elastic strain energy.
Another one-dimensional approach utilizing basic relations such as the
mass and momentum balances with some constitutive equations has been
proposed a number of times for calculating the tensile thread parameters.
Such an approach was used in Refs 77 and 439, for example, where the
calculation of the thread parameters was only carried out downstream of
the point corresponding the maximum swelling ratio. This model, which
ignores the polymer deformations in the upstream region of capillary
flow, can also be treated as semi-empirical, since it requires information
about the cross-sectional area at the point of maximum swelling; a
parameter that is sought for itself.
The main objective of this section is to develop a theoretical model for
solving the problem in question without the requirement of any empirical
data other than the rheological parameters in constitutive equations. The
parameters should be determined in visco metric tests separately.

10.7.1 Formulation of the Problem

We will study the steady, isothermal and inertialess flow of a viscoelastic


liquid flowing out of a capillary die of length L and radius Ro (Ro/L4, 1).
The flow, sketched in Fig. 4.17, is driven by the pressure Po in the
pre-entry capillary region and by a stretching force F acting in axial
z-direction on the end of a liquid extrudate. It is sometimes convenient to
use instead of F, the average tensile stress (Jo=F/nR'5 at the die outlet,

328

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

and instead of flow rate Q, the average speed of the liquid


<v z == u = Q/nR6, extruded from a circular die. It is also convenient to
consider the space occupied by flowing polymer as being divided into
three regions: region I situated inside capillary, with a small transition
region II near the die exit, and region III outside capillary, with the
free-surface flow of extrudate. Within region I, the flow of polymer can be
treated as steady and completely developed provided that the entrance
effects can be neglected. Within region III, which starts just beyond the
die outlet, we shall consider the steady flow of polymer to be only slightly
inhomogeneous to exploit the one-dimensional propositions for constitutive equations and averaged conservation laws. To avoid extremely
complicated analysis of flow in transition region II, we exploit the idea of
Refs 437 and 438, using the equating of elastic energy fluxes for regions I
and III as a matching condition.
In both regions I and III we shall use the same, most simplified
constitutive equations:

>

detck= 1

(k= 1, 2)

W == pf = (t) L I-lk tr( Ck - 15)

(10.61)

All the notations in eqns (10.61) coincide with those in eqns (1.71)-(1.76),
specified with nk = 1, s = 0, N = 2, )\ = O. In the case of simple shear, these
equations are presented in detail as eqns (5.13)-(5.14) with h =0, and for
simple elongation as eqns (6.17).

10.7.2 Steady Capillary Flow in Region I


Steady capillary flow has been considered in Section 8.4.2 for an arbitrary
number of relaxation modes N, and represented by dimensionless formulae (8.26)-(8.30). In the particular case of the two-modal approach
(10.61), it is convenient to scale the shear stress by the lowest modulus 1-l1
and shear rate by the highest relaxation time 8 1 . By introducing the two
dimensionless parameters
(P~1, v~1)

we can rewrite eqns (8.28)-(8.30) in the new dimensionless form:


2r
2vpr
2
0"12 = 1 +J1 +4r + 1 +J1 +4p 2 r

2;

(10.62)

329

Flows Close to Simple Shear and Simple Extension

W(r) =(2 + 2J1 + 4r2)1/2 + v(2 + 2J1 + 4p2r2)1/2 - 2(1 + v)


v(r)=ln(1 +J1 +4x 2

(10.63)

m +(vIP)ln(1 +J1 +4p x m


w

cp(r) =2(l12(r)[(1 + J1 +4r 2)-1 IJ1 + 4r 2

+vp(1 +J1 +4b 2 r 2 )-1/J1 +4p2r2)


where r wand ,ware the dimensionless shear rate and shear stress at the
capillary wall. Equation (10.63) are employed below to calculate the
values U(r w) and <VW)in(rw).
It should be noted that the expression for shear stress (l12 (r) in eqns
(10.63) has the limited value 1 + v as r ~ 00. Therefore we assume that
'w < 1 + v. Otherwise one should use more complicated formulae (1.71)
either with N> 3 or with nk> 1 or both.
Concluding this section, we should note that, in practice, the parameters Po and (lo, included in the expression for the pressure gradient P
(see eqns (8.26) and (8.27)), usually satisfy the inequality: Po~(lo. In this
case, stretching the extrudate has no effect on the flow of polymer in
capillary die. However, for extremely large stretching forces, the flowing
polymer can be detached from the capillary walls,436 which may occur if
the total normal stress at the wall, (lm exceeds the polymer-to-wall
adhesion stress (lad. Even if we know the value (lad from independent
experiments, the precise estimations of the critical stretching force is
impossible in this model because eqns (8.26)-(8.30) do not take into
account the complicated phenomena of flow restructuring in the die exit
region II.
10.7.3 Stretching Polymer Extrudate Outside the Circular Die
Outside the die outlet (region III) we assume the steady inertialess
polymer flow in the stretched extrudate to be only slightly inhomogeneous (ldR/dzl ~ 1). Under steady conditions, the values of
stretching force F and flow rate Q remain constant along the jet. This

330

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

results in the common relation O"(z)/v(z) = F/Q = constant. It is convenient


to introduce the dimensionless variables and parameters:
0"* = 0"//110

v*=vedR o,

z* =z/Ro

O"~=O"O//11o
t* = t/e b

R*=R/Ro

(10.64)

Then using eqns (10.53)-(10.57) of Section 10.5, which establish the


correspondence between homogeneous and inhomogeneous extensions,
we can write the set of equations for stretching the extrudate jet in the
dimensionless form (see also eqns (6.17)):
(6/x) dx/dt+(x 3 -l)(x+ 1)/x2 = (6/s) ds/dt
f3(6/y) dy/dt+(y3 -l)(y+ 1)/y2 = f3(6/s) ds/dt

(10.65)

0"0s=x 2 _x- 1 +v(Y2 _ y-l)

z=uj;S(tddt b

s=R- 2

(10.66)

where s(t) is the total strain (Hencky measure). Also, to simplify the
notation, the stars in dimensionless variables in sets (10.65)-(10.66) are
omitted. The closed set of equations (10.65) should be solved under the
initial conditions:
(10.67)

where the parameters Xo and Yo are the subjects of determination. To


evaluate these we use the dimensionless expression for the elastic energy
in extrudate flow:
W(t)=x 2 +2X-1-3 +V(y2 +2y-l_3)

(10.68)

Now using the fact that for weakly inhomogeneous extension


<W(t)U(t)

W(t)U

we can write down the matching condition for inlet and outlet flow of
elastic liquid as follows:
(10.69)

Additionally, the condition of continuity of averaged (over cross-section)


longitudinal stress at the die outlet gives:
0"0 =x6

-xo 1 +V(Y6- Yo 1)

(10.70)

Flows Close to Simple Shear and Simple Extension

331

The matching conditions (10.69) and (10.70) whose left-hand sides are
known from capillary flow, deliver two additional equations to determine
the unknown values Xo and Yo.
The set of algebraic equations (10.69)-(10.70) is easily decomposed into
two independent equations
(a-xo 1 )2(b -x6)= v 3 ,

Yo = vXo/(axo -1)

a=1+v-1(0"0-<vW);nIU),

b=a+O"o

(10.71)

which readily allow the calculation of the parameters Xo and Yo. As soon
as the set (10.65) with conditions (10.71) is solved, the dimensionless
dependence R(z) is restored using parametric relations (10.66)
In the particular case 0"0 = 0, the problem is reduced to computing the
swelling of free polymer extrudate exiting from the capillary. VJY often
the simpler problem of determining the swelling ratio '}' = II 8( co) is
considered. This magnitude can be immediately calculated by using the
formula (6.45) with a = 1 and the dependence y(x) being established from
the third equation (10.65) at 0"0=0.
As a comparative example, we also consider the case given by the
solution of eqns (10.65) with the initial conditions:

Xl t =0=80,

ylt=o =80

(10.72)

which models the stretching of polymer extrudate, ignoring the polymer


flow inside capillary. Here the parameter 80 being an elastic deformation
'jump' at z = 0 (t = 0), is determined from the third eqn (10.65) at
8=X= y=80:

86 -860"0/(1 + v)-l =0

(10.73)

Proposed in such a way, the problem coincides with that considered in


Section 6.5 for the homogeneous extension of a viscoelastic cylinder from
the rest state under action of a constant force.
All the problems formulated require numerical computations. Some
results of these for certain values of the parameters are compared with
experiments in the next section.
10.7.4 Experimental Data and Their Comparison with
Theoretical Calculations
In the experimental studies, the PIB P-20 with molecular mass 10 5 and
density p = 910 kg/m 3 was used. This polymer was slightly different from

that investigated in Chapters 5 and 6.

332

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The stretching of the polymer extruded from the capillary die as well
as its free swelling, were studied in the apparatus shown in Fig. 4.17.
Here, under constant pressure Po, the liquid polymer jet (2) was extruded from the capillary (1) of length L= 10 cm and radius
Ro = 0.25 cm. In stretching experiments, with capillary flow also maintained under given pressure, the end of the polymer extrudate was
attached to the bushing (3) connected to the thread (4) passed over the
roller (5) to the load (6). In order to remove the effect of gravity and for
thermostatting, the polymer was extruded into the water bath (7). The
extrudate profile was photographed on a 1: 1 scale with a camera (8)
movable in the axial direction. A decoder was also used to enlarge the
dimensions of the extrudate profile. The typical photograph of the
extrudate is shown in Fig. 10.15. In order to determine the limiting free
swelling ratio y at F = 0, the specimen's diameter was measured at a
downstream position where the diameter was practically constant. A
typical time of motion of the extrudate's cross-section from the die
outlet to the measuring position was approximately 10 min. At a given
value of capillary pressure Po, the flow rate Q was determined by
weighting of polymer jet pieces extruded in a certain time. Thus, the
quantities R(z), y(Po) and Q(po) were measured in the experiments. For
all the experimental data, the errors did not exceed 7%.

Fig. 10.15. Photograph of a polymer jet extruded from a capillary and extended under a constant stretching force.

The rheological properties of the PIB were initially studied in the


Rheogoniometer R-18 at a temperature of 200.5 0c. It was found that
the maximum Newtonian viscosity of the polymer 1] ~ 1.3 x 106 Pa s - 1
and its high elastic modulus Ge~4.4 x 103 Pa. The open circles in

333

Flows Close to Simple Shear and Simple Extension

fOf

Fig. 10.16. Flow curve for PIB P-20. Data are denoted by points; the line is
drawn according to the expression for 0"12(1) in eqn (10.63).433

Fig. 10.16 demonstrate the flow curve of the polymer. The line is the
theoretical flow curve according to dependence 0"12(r) in eqn (10.63)
with the following values of fitting parameters: 111 =2.7 x 103 Pa,
112 = 7.8 X 104 Pa, 8 1 = 380 s, 82 = 3.5 s. These parameters were determined from the experimental flow curve and the empirical values of
parameter 1] and Ge given above, by using the linear relations (1.91):

Ge=1]2/~1]k8k
One can see that the basic inequalities (1.69) are satisfied. The values of
dimensionless parameters f3 and v are: f3=9.21 x 10-3, v32.6. For
y< 10 - 3S -1, the flow curve has a Newtonian region which is not shown
in Fig. 10.16. For y> 10- 1 S - 1, the description of the flow curve by the
expression 0"12(r) given in eqn (10.63) becomes impossible.
The results concerning the free swelling are demonstrated in Fig.
10.17(b) where the 'capillary flow curve' Q/nR'6 versus Tw is also shown. It
is seen that the discrepancy between the experimental flow curve and its
theoretical description displayed in Fig. 10.16, also results in the appearance of that in the capillary flow curve (line and symbols 1, respectively).
Symbols 2 and the corresponding solid line in Fig. 1O.17(b) represent the
experimental data and calculated results, based on eqn (6.45), for the
values of swelling ratio y versus Tw' As seen, these dependences are
invariant with respect to value Ro and L, and increase monotonically
with the increase in Tw'

334

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

a)

ZP

1.5

1.0

a5~------~--~--~~----~S
{04
fa

CCw

(Pa)

Fig. 10.17. Free swelling of polymer jet extruded from capillary. (a) Sketch of
the polymer flow. (b) Plots of qjnR5 (0, curve 1) and the limiting swelling ratio
)I(i:,., curve 2) versus shear stress at the capillary wall 'w' Solid lines denote the
appropriate theoretical plots. 433

Figure 10.18 and 10.19 compare the experimental and theoretical


plots of dimensionless radius of extrudate R versus the dimensionless distance from the die outlet z for the values of dimensionless
capillary wall shear stress 'w=7.47, 18.5 and 23.15. The relevant points
and lines are denoted by 1, 2 and 3. Figure 10.18 corresponds to free
swelling and Fig. 10.19 to swelling and stretching under constant axial
force F with the respective dimensionless initial extension stress
0"0=2.59.
The experimental profiles R(z) were obtained by means of photographing the steady polymer jets extruded from the capillary. In the
case of free swelling (F = 0), the dependences R(z) increased down-

335

Flows Close to Simple Shear and Simple Extension

2
0

*Q::

01
.2
o 3

0.25

0.50

0.75

1.00

1.25

z*/<v*>

Fig. 10.18. Dimensionless profiles R(z) for free extrudate swelling. Data denoted
by symbols 1, 2 and 3 correspond to the values of dimensionless wall shear stress
"rw = 7.47, 18.5 and 23.15, respectively. The solid lines 1, 2 and 3 denote relevant
theoretical dependences. 43 3

1.6

D 1
2
o 3

..

Q::

1.0
0.8
0.6
0.4
0

"-

'-...

'-...

........

-- ----

0.25

0.50

3
2

0---0-1

---- --- ---

0.75

1.00

1.25

1.50

z*/<v*>

Fig. 10.19. Dimensionless profiles R(z) for the stretched polymer jet extruded
from the capillary die (ao = 2.59). Data denoted by symbols 1, 2 and 3 correspond
to the values of dimensionless wall shear stress "rw = 7.47, 18.5 and 23.15,
respectively. The solid lines 1, 2 and 3 denote relevant theoretical dependences.
The dashed line indicates the dimensionless theoretical profile without taking into
account shearing polymer in capillary, i.e. the solution of eqns (10.65HlO.66) with
initial conditions (10.72) and 80 = 1.026 366. 433

336

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

stream monotonically; their limiting values corresponding to the swelling ratios}, shown in Fig. 1O.17(b) by symbols 2 (open triangles).
Under a stretching force the plots of profile R(z) passed through the
maxima at approximately the same dimensionless value Zm. The maximum value Rm is larger, the more intense is shearing inside the capillary,
i.e. the larger is the value Tw. The occurrence of swelling under stretching
force in extrudate implies the initial macromolecular disorientation when
the polymer exits the capillary, which changes to the increase in orientation when the polymer moves further downstream.
The theoretical calculations represented in the figures were accomplished using eqns (10.65)-(10.66) with initial conditions (10.67) fixed by
eqns (10.71). All the computational parameters are presented in Table
10.2. This table shows that the values of dimensionless shear rate at
capillary wall, r w, the dimensioness flow rate U, the dimensionless elastic
energy flux <vW), and the parameter xo(xo> 1) increase monotonically
with the growth of T w , whereas the parameter Yo(Yo < 1) decreases with
increasing Tw. It is of interest to note that at (J 0 = 2.59 the values of Xo are
higher and those of Yo are smaller than at (J 0 = o. Calculating the
functions x(t), y(t) and e(t) revealed the following behaviour of these. At
(Jo =0, the functions x(t) and e(t) decrease while y(t) increases monotonically. At (Jo =2.59, the behaviour of all the functions is more complicated.
Namely, x(t) decreases initially, and increases again after passing through
a minimum; the function e(t) behaves similarly but with the minimum
attained earlier relative to that of x(t), and the function y(t) still monotonically decreasing. All these facts are quite conceivable from the
viewpoint of the mechanical behaviour of the rheological model under
consideration.
The dashed line in Fig. 10.19 corresponds to the dimensionless plot
R(z) calculated at (Jo = 2.59 without an allowance for capillary shearing,
i.e. by using the initial conditions (10.72) with the value eo = 1.026366 as
the solution of eqn (10.73).
Table to.2 Parameters of computations
No. of

'Lw

rw

(v)

(VW)in

regime

1
2
3

Xo

(Jo=O
7.47
18.5
23.2

22.5
82.0
137

5.29
17.8
26.8

25.3
229
477

2.57
3.66
4.14

Yo

(Jo=2.59 (Jo =0 (Jo = 2.59


2.62
3.74
4.22

0.937
0.867
0.830

0.960
0.887
0.848

Flows Close to Simple Shear and Simple Extension

337

In the experiments, the value of (Jo did not exceed 0.7% of the
minimum capillary pressure value. Therefore the stretching of the extrudate had no effect on the flow rate. This also explains why in the
experiments, the polymer did not break away from the capillary wall.
The latter was justified by the fact that the diameter of extrudate just
beyond the die outlet coincided with the internal diameter of the
capillary.
The theoretical results reveal at least a semi-quantitative agreement
with the data, describing all their features. However, the theoretical
values regularly exceed those of the experimental data. Maximum discrepancies of 13.8% occur in regime II for free extrudate swelling. In the
case of the stretched polymer extrudate, the discrepancies are smaller
with a maximum of 7.7% in regime III.
Apart from the possible inadequacy of the constitutive equations used
within the shear rates under study (or a non-optimal choice of the parameters in these), the following may lead to errors in the proposed
theoretical model:
(i) dissipative losses within region II (see Fig. 10. 17(a)) are not
negligible;
(ii) an average description of the flow in polymer extrudate is not
acceptable near the die outlet.
The fact that all the theoretical curves lie above the relevant experimental ones gives support to (i). The fact that the disagreements are
smaller for the stretched polymer extrudate gives support to (ii), since
by virtue of evidence given in Ref. 434 one can assert that the velocity
distribution in the stretched polymer extrudate at the die outlet is
more homogeneous than in the case of free swelling extrudate. Thus it
seems that both sources of errors arise in the theoretical model. However, although the dissipative losses can be roughly estimated, it is
impossible within the framework of the theoretical model proposed, to
take into account the inhomogeneity in the initial region of the polymer
jet.
It is also worth mentioning that the matching condition (10.69) leads to
the closed set of equations only in the case of two relaxation modes
(N = 2). In the case of arbitrary N, one can use the same matching
conditions (10.69) for every kth relaxation mode with the partial free
energy function Wb taking into account that immediately after the die
outlet there is an elastic jump 80.

338

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

10.8 DRAWING POLYMERIC FLUIDS FROM A RESERVOIR


WITH A FREE SURFACE: OPEN CHANNEL (ELASTIC) SIPHON
10.8.1 Basic Effects
These phenomena hold for some polymer solutions and are usually
attributed to the spinnability of polymeric liquids. In Ref. 22 where they
were studied first, as well as in Ref. 198, a tube with a negative pressure
was lowered into a polymer liquid (usually a few per cent aqueous
solution of polyethylene oxide, PEO), contained in a reservoir, and the
liquid was sucked from the reservoir into the tube. Then raising the tube
with continued sucking resulted in the formation of a long free jet of
elastic liquid between the raised tube and the reservoir. This is schematically shown in Fig. 10.20(a). A similar phenomenon, the open channel
(elastic) siphon 440 is sketched in Fig. 10.20(b). The vessel (1) containing an
elastic liquid was tilted so that the liquid was poured out over the rim.
Then the vessel was returned into the initial, vertical position, while the
solution continued to pull itself from the vessel under the action of the
external part of the jet.

G)

Fig. 10.20. Schemes of drawing of polymeric solution from a reservoir with a


free surface: (a) suction into a capillary; (b) open-channel (elastic) siphon.

Another experimental scheme for drawing polymer solutions from a


vessel with a free surface by a rotating drum 294 is discussed in Chapter 4
and represented in Fig. 4.15. The PEO with the Newtonian viscosity
YJ ~ 0.1 Pa s - 1 was used in those experiments, where the length of stationary drawn viscoelastic jets could reach half a metre. It was found that the
steady axisymmetric flow in the jet (Fig. 1O.21(a)) was possible in a
narrow region of flow rate q. At small flow rate, the jet thrown on the
rotating drum was broken, while at larger flow rate, it lost the axisymmetric form (Fig. 1O.21(b)); the diameter of jet varying with time

Flows Close to Simple Shear and Simple Extension

339

Fig. 10.21. (a) Symmetrical and (b) asymmetrical forms of jets for a 0.5%
aqueous solution of PEO drawn from a reservoir with a free surface.

sporadically. In the experiments,294 the formally defined ratio of stress to


the strain rate, commonly treated as elongation viscosity, exceeded the
initial Newtonian viscosity by lO4 times. It is also worth noticing that
neither such a Newtonian liquid as honey, nor such viscoelastic fluids as
solutions of butyl rubber in transformer oil (whose rheological behaviour
has been discussed in Chapter 5), both having the Newtonian viscosity
'1 ::d0 3 Pa s -1, could be drawn by the method.
10.8.2 Theoretical Model
A successful theoretical model 441 has recently been proposed to describe
the experiments,294 where the deformation of free jet in region 1, just
above the bottom interface, was assumed to be pure elastic. This can be
treated as a relaxation transition from the initial, almost Newtonian flow
in region 3 under the interface, to the elastic solid deformation in the jet
region, with the fluidity loss in an intermediate zone 2 (Fig. 4.15, lO.21).

340

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Within region 1, the steady equations of inertialess momentum and


mass balances, averaged over the jet cross-section, are of the form:
d/dz(I1S - 21'fo) = pgS

Sv = q = constant
(10.74)

where S is the cross-sectional area of the jet of actual radius r(z), 11 is the
stress, p is the density, l' is the surface tension coeffient, g is the
acceleration due to gravity, and z denotes the coordinate of the jet's axis
counted off from the free surface level of the liquid in the container.
Inertia terms are neglected in the first, momentum balance equation in
eqns (10.74).
Because of deformative orientation, the rheological behaviour of the
solution in region 1 is assumed to be identical to that of a crosslinked
rubber. Then, using the elastic limit for eqns (6.16), we can write the
expression for stress 11 as:
(A~

1)

(10.75)

where !1 is the elastic modulus and n is a constant (1 < n < 2). Assming that
z ~ R, where R, is the meridianal radius of 'viscoelastic meniscus' we
obtain in the sightly non-uniform elastic approximation that
.1= Ao(ro/r)2 =AoSo/S

(10.76)

Then the solution of eqns (10.74)-(10.76) is given by:


Go(l-l/n)[(So/st -lJ -1' jn/So(jS/So -1)= pg(z-R)

11= Go(So/st-1'F/S,

Go = 110 +1'jn/So

(10.77)

where .1 0 , ro, So and 110 are the elastic strain, jet's radius, cross-sectional
area and stress at the level z = R, which should be determined by
matching the behaviour of the liquid in regions 2 and 3.
As shown experimentally,294 the flow in region 3 (z < 0), is not
important, because the stress in the jet is determined mainly by the weight
of liquid column over the free surface (z > 0) and, to a certain extent, by
the influence of surface tension. The values of ro, Rand 110 were estimated
in Ref. 294 by using dimensional and geometrical arguments as follows:
R=rJ.(qB)1/3;
v=rJ./fJ;

ro=fJ(qB)1/3;
110~pgmR+2v1'/R
2
m=v [i+(1 + 1M2-n(l + 1/v)/2]

(10.78)

where B is the relaxation time, rJ. and fJ are two numerical parameters
which were estimated in Ref. 294 by the fitting of experimental and

341

Flows Close to Simple Shear and Simple Extension

theoretical jet profiles. The first term in the expression for the stress a in
eqn (10.78) arises because of the weight of the liquid column and the
second one describes the contribution of surface tension. According to the
expression for ao as a function of R, there is a minimum value, min ao,
with corresponding values of Ro and qo, given by:
minao = J8pgymv,

Ro = J2vy/(pgm),

8qo = [2vy/(pgmC(2)] 3/2

(10.79)

The values of Ro and qo are treated as the critical ones, below which the
steady flow does not exist.
When the effect of surface tension in region 1 is negligible, the solution
in this region is represented as:
r/ro = [z/(cR) + l-l/c] -1/2n,

v = q(z/cR)l/n/(nr'f;),

a = pg[nz/(n -1) + Rm(1-1/c)]

K=v/(nz)

(10.80)

(c = m(l-l/n))

It is seen that when z ~ R, the stress is independent of the flow rate and is
related only to the action of gravity force.
In experimental works, as well as in industry, the ratio 1'/* = a/K is often
used as a measure of elongational viscosity. When z ~ R, the expression
for 1'/* is:
2

1'/* = nropgz (cR/z)l/n


q(n-1)

(10.81)

which grows unboundedly with increasing z.


10.8.3 Comparison with Experiments
In experiments,294 the following values of parameters were found for
0.5% aqueous PEO solution:
p~O.l
Jl~2

Pas 2 cm- 2 ,

Pa,

n~

1.34,

y ~ 6.4 Pa cm -1,
C(~0.87,

8 ~ 0.7 s -

v=C(/f3~2.6

(10.82)

With these values, the lower critical values of qo and Ro are:


qo ~ 0.0932 cm 3 s -1 and Ro ~ 0.341 cm. This is in very good agreement
with experimental values: qo ~ 0.093, Ro ~ 0.35 cm. In the experiments, it
was impossible to withdraw a jet of the polymeric solution for q<qo (or
R < Ro). The experiments also revealed some upper critical values, q* and
R* (q*~1.14cm3s-1, R*~0.8cm), related to unstable non-symmetric
disturbances.

342

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Figure 10.22 and 10.23 demonstrate the comparison between data and
calculations for the jet's profile for various flow rates and stress distributions along the jet, respectively. Overall, the discrepancy between the
results of the calculations and the data does not exceed 20%. Note that
all the calculated profiles are somewhat below the corresponding experimental curves.

o1
x2
'13
4
~5

V' V'

16

24

z (em)

Fig. 10.22. Theoretical (solid lines) and experimental (symbols) dependences of


the jet radius on the longitudinal coordinate z for a 0.5% aqueous solution of
PE~. Symbols 1-5 correspond to q= 1.14, 0.58, 0.36, 0.17, and 0.09 cm 3 s- 1,
respectively.441

10.8.4 Generalizations
Let us now consider the situation shown in Fig. 10.24, in which the jet
(1) with length Lb passes over the two wires (2) with horizontal flow (3)
between them, and then falls under its own weight as the liquid column
(4) with length L 2 This problem has a steady solution, if a viscoelastic liquid from a class considered in Section 10.8.3, is kept in both
reservoirs 5 and 6 at certain constant levels. To construct the solution of
the problem we can apply the results of the previous problem studied
in Section 10.8.3. If the lengths of both jets are measured from the
free surface levels, then at some flow rate q, defined by certain conditions, both jets will have identical profiles and equal lengths. This
result follows immediately from eqns (10.76), (10.77) and the above

343

Flows Close to Simple Shear and Simple Extension

60
3
ru
a.

50
2

'$2 40

)(

30

20
10
0

10

20

z(cm)

Fig. 10.23. Plot of stress (J versus


longitudinal coordinate z for flow rate
q = 0.38 cm 3 s -1. The solid line indicates the calculated dependence; the
points represent the experimental
results. 441

Fig. 10.24. A possible design of openchannel siphon. 441 The jet (1) passes
over the two wires (2) with horizontal
flow (3) between them, and the falls
under its own weight as a liquid column (4).

dimensionality considerations. Here the length L2 is defined by the


relation:
(10.83)
Here Ll is given and the functions r(z) and (J(z) are defined by eqn (10.80).
If the frictional force Fcr of the jet on the wires (2) is negligible, then
L2 ~ L 1 It is of interest to note that in the case of an open-channel
siphon, the free-falling jet (4) widens as it falls. Such behaviour is related
to the rubber-like elasticity. When the siphon cannot occur, the jet of an
elastic liquid is, as usual, convergent.
10.9 PROPAGATION OF NONLINEAR WAVES ALONG A
VISCOELASTIC BAR
This is a classical dynamic problem which was studied in Ref. 442 within
the nonlinear 'bar' approximation443.444 developed for elastic solids. A
number of general results in nonlinear dynamic elasticity and solutions of
various problems are presented in Ref. 445. Relationships at strong and

344

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

weak discontinuities in hereditary media of a general type were studied in


Ref. 446. We will employ below the viscoelastic Maxwell-like constitutive
model discussed in detail in Chapter 1.
10.9.1 Fundamental Equations

In the following, we consider the bar motion to be close to uniaxial


compression-extension and use the description averaged over its crosssection. It is assumed that apart from specific cross-sections, all quantities
vary little along the bar and that the length of the bar is much greater
than its cross-section The bar is assumed to be incompressible with a
density p.
The equations of mass, momentum and energy balances in the bar
approximation are of the form:
oJ +oxUv)=O,
Ot[pf(U +v 2 /2)] +ox[pfv(U +v 2 /2)] =oxUv(J-q)-a(T - To)jJ

(10.84)
where f is the cross-sectional area of the bar, v is the velocity, (J is the
normal stress (defined as in the uniform case, with the zero stress
condition on the bar free surface), U is the specific internal energy, q is the
longitudinal heat flux, T is the temperature, To is the ambient temperature, a is the lateral heat exchange coefficient, x is the longitudinal
coordinate, and t is the time; the quantities f, v, (J, U and T being
considered as averaged over the bar's cross-section.
To close the set (10.84) it is necessary to formulate the thermodynamic
and rheological constitutive equations. Let us assume that a bar consists
of a dissipative rubber-like material of the raw (uncured) rubber or
polymer melt type. As discussed in Chapter 1, the following rubber-like
thermodynamic functions
U = U o +c;.(T - To),

S=So -/l'P(A)+c;.ln(T/To)

F = F 0 +(c;. -co)(T - To)-c;. T .In(T/To )+ ,a'T'J'(A)/p

(10.85)

are valid in the one-dimensional approach employed. Here A is the


recovery strain, ,a'Tis proportional to Young's modulus E, and c,\ is the
specific heat capacity at constant A, which is assumed to be constant over
a broad enough range of temperatures. For compression, O<A< 1, and
for extension, A> 1.

345

Flows Close to Simple Shear and Simple Extension

The dimensionless function 'P(2) in eqn (10.85) is a certain elastic


potential. As a good example, we can present BST elastic potential S7
'P(2) = 2I/1n(2) + b[ 1/1n(2)]m,
(1 <n<4,

b~O,

(10.86)

m>O)

valid in a large region of 2 variations. The case n = 2, b = 0 corresponds to


the classical potential of rubber elasticity. In Fig. 10.25 solid lines
represent the BST elastic potential, and dashed lines, the classical
rubber-like potential.

i..

'V'"

@
A

Fig. 10.25. Qualitative forms of: (a) the dimensionless elastic potential '1'(.1.); and
(bHd) its three derivatives. Solid lines represent the BST potential (eqn (10.86)),
dashed lines represent the classic potential (n = 2, b = 0).442

In the case of slightly non-uniform compression-extension, the thermorheological constitutive equations can be written in the form:
Ot 2 -1 + ox(v2 -1) = cp(2)j[68(T)2],
0"(2, T)=Il'TA'P'(2),

2'P s (2)='P(2)+'P(2- 1 ),

8(T)= 8oexp(AEjRT)
cp(2) =

-p- 1 djd2exp[ -fJ'Ps(2)]

q= -K(2, T)oxT

(K>O)
(10.87)

where 8(T) is the relaxation time and K is the coefficient of


thermal conductivity; the dissipative term cp(2) describing the hardening

346

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

phenomena with numerical parameter /3, /3 E (0, 1). When investigating


highly intense waves, the longitudinal thermal conductivity, as well as the
possibe Newtonian viscous terms, are omitted because these are small
outside some very narrow regions, and the surfaces of weak and strong
discontinuities will later be used as a substitution for the regions where
these values are very large. Hence, it is further assumed that q == O.
From eqn (10.84) it is possible to obtain an equation for balance of
specific internal energy U which, by considering the expression for U
from eqn (10.85), has the form of temperature balance:
(1O.88a)
If the heat capacity at constant stress,

Ca ,

is introduced by common

formulae:

Jl'- )
( y -pc).

(10.89)

eqn (10.88a) can be rewritten in the dual form:


pC a [ot(fT) + ox(fvT)] = ofcp(2)/[68(T)] -rx(T - To)J1 (1O.88b)

For the polymer materials in question, the value of y is very small


The first term on the right-hand side of eqn (10.88a) is the
power density which, generally speaking, is not positive definite. The first
term on the right-hand side of eqn (10.88b) is disspation and always
positive.
If 8- 00 (/3 . . H:tJ ), the equations in this and following sections describe
the motion of incompressible perfectly elastic rubber in the 'bar' approximation.
(y~ lO- Z).

10.9.2 Weak Discontinuities and Shock Waves in a Viscoelastic Bar


Let x*(t) be a curve in the x ....t plane where derivatives of continuous
functions gk(X, t) have discontinuities. It is known that the characteristic
roots for a hyperbolic set of equations coincide with weak discontinuity
propagation velocities x~(t). For the closed set of eqns (10.84), (10.85) and
(10.87) with condition q = 0, the characteristic roots have the form:

rxl.Z=VU..
rx3=V
2
u; =ui + a /(pc;. T),
pui = 22d/d2(a/2)IT

(10.90)

where Us and UT are the adiabatic and isothermal velocities of sound in


deformed medium, respectively. It is seen that Us ~ UT; the equality being

Flows Close to Simple Shear and Simple Extension

347

,2!!!Y for nondeformed bar (,.1,= 1, i.e. a=O), where Us=UT=


-J E/p(E = 3nJ-l'T). In accordance with the expression for a(A) in
(11.23), uf = p - 1 J-l' n 2'P"(A). Thus the inequality 'P"(A) > 0, which is

achieved
uo(T) =

eqn
the condition of thermodynamic stability in the one-dimensional case
under study, also provides the condition of strong hyperbolicity or
Hadamard stability (uf >0) for the set of equations in question. It is also
possible to obtain the equation relating the values of dynamic variables a
a weak discontiuity x* (t).
Strong discontinuities can also exist in the set (10.84), (10.85) and
(10.87), which is reminiscent of the gas dynamic equations, with the
cross-sectional area f corresponding to the gas density. Let xo(t) be a
line of strong discontinuity in the x-t plane. Choosing the x-axis
direction so that Xo > 0, we denote all values in front of the shock wave
(x=xo-O) by the superscript 1, and behind the shock wave (x=xo+O)
by the superscript 2. The conditions at jumps for the above set are:
[f(v-xo)] =0,
[A-1(V-XO)] =0,

p[f(v- XO)2] = [a!]


p[f(v-xo)(U +v2/2)] = [fva]

(10.91)

Here the standard notations for the jump are used: [y] = Y2 - Yl.
The first to fourth equalities in eqns (10.91) correspond to the conservation of mass, momentum, elastic strain and energy fluxes, respectively, through a shock wave; the mass flux being negative in a fixed
coordinate system considered here. Using eqns (10.91), it is easy to find
that
[fA] =0,

p[v] = - [A]J[a/A]/[A]

(10.92)

The first equality in eqns (10.92) shows that the passage through the
shock wave is accompanied by the pure elastic deformation; the second
one gives the alternative:
[a/A] <0,

[A] <0,

[v]>O

(1O.93a)

[a/A] >0,

[A] >0,

[v]<O

(1O.93b)

Inequalities (1O.93a), which characterize the compression shock waves


usual for gas dynamics, can occur for the viscoelastic liquid in the
region of compression (A < 1) and moderate extensions. Unlike gas
dynamics, at very large extensions, the extension shock waves can occur
in rubber-like materials under study, which are characterized by ineqns
(10.93b).442

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

348

Using eqns (10.91) and thermodynamic relations (10.85), one can


obtain the expressions for Hugoniot's adiabate, entropy and temperature jumps at the shock wave, represented, respectively, as follows:
1(2 + y[A]\f'1)/(A1 \f'd

(10.94)

T 2(2 -Y[A]\f 2)/(A2 \f2) = T 1(2 + y[A]\f'1)/(A1 \f'd

(10.95)

(J2(2 - Y[A]\f2)/(A2 \f2) =

[S]
c...

(J

In 2+Y[A]\f'1 -y[\f]

(10.96)

2-y[A]\f2

Due to the dissipation associated with the shock waves, [S] >0. This
allows us to establish the conditions for the existence of the shock
waves. When \f'" changes sign, as in the case of the rubber-like elasticity
shown in Fig. 10.25, it is also necessary to consider their stability (see,
e.g. Ref. 447). These conditions are of the form:
sgn( 8[ S]/8A2) = sgn[A],

sgn(8[S]/8A 1) = -sgn[A]

(10.97)

The condition of existence, [S] >0, and stability conditions (10.97) made
it possible to establish in Ref. 442 a complete diagram of stable shock
waves propagating along a viscoelastic bar, and to consider some useful
examples.
10.9.3 Striking a Viscoelastic Bar of Finite Length
Against a Rigid Obstacle

This problem was studied numerically in Ref. 442 in the following


formulation. Prior to impact, the bar is considered as homogeneous,
either non-deformed or uniformly extended, having the velocity Vo. The
coordinate origin, x = 0, is chosen at the free bar end at the moment of
its contact to the wall. The problem was studied in isothermal approximation, which is valid when the temperature jump at the impact is
negligible. This takes place if the relaxation time is small enough. In this
approach, the closed set of equations consists of mass and momentum
balance (the first two equations of eqns (10.84 and the rheological eqns
(10.87) with q = 0 and BST potential (10.86). The initial and boundary
conditions are formulated as follows:
t=O:

v = vo,

t>O:

Vlx=L=O,

f = fo = constant,

Alx=l(t)= 1

(O~x~L)

(10.97)

where l(t) is the free bar end coordinate. The boundary condition A= 1
at x = l(t) in eqn (10.97) corresponds to the dynamic condition (J = 0 at
the free bar end.

349

Flows Close to Simple Shear and Simple Extension

Now we discuss the phenomena that occur when a viscoelastic bar


strikes a rigid obstacle. Just after impact, a compression wave propagates along the bar, which after reaching the free bar end, is then
reflected in the form of an unloading wave travelling towards the rigid
obstacle. The wave of unloading, on reaching the obstacle, is reflected
from it and transformed into an expansion wave.
During this time the relaxation processes occur in the viscoelastic bar,
so that the maximum stress values at the point of contact with the wall,
decrease with every cycle. At the same time, beginning with the moment
of initial impact, bar-wall adhesion forces appear. The adhesion may
considerably change the character of the bar motion, beginning with the
moment when the extension wave first appears. If at this moment, the
adhesion forces exceed the extending force in the bar at the contact
point, the bar will 'stick' to the wall and its further motion will be
governed by the decay of the oscillations due to the relaxation phenomena. When the extending force exceeds the adhesion ones, the bar will
be detached from the wall with a certain mass average velocity VI. Due
to the relaxation processes it will be accompanied by a loss of kinetic
energy in the bar.
The bar-wall separation conditions are modelled III Ref. 442 as
follows:

a(L, t* + 0) > kE,


a(L, t* +O)<kE,

v(L, t* + 0) < 0

v(L,t* +0)=0

(1O.98a)

(Vlx=L=O, t>t*) (1O.98b)

Here, the adhesion stress ao was roughly taken as kE, with a numerical
coefficient k of the order of unit. The condition (10.98a) corresponds to
detachment from the wall, and (1O.98b) is the stick condition.
Employing as characteristic values of dimensionality of length 2L,
time 2L/vo, cross-sectional area /0, and stress 2jl, reduces the above set
of equations to the dimensionless form, with the three dimensionless
parameters:
{J

(10.99)

where Mo is the initial Mach number.


The calculations have been conducted in a movable coordinate system connected to the movable end of the bar, on the basis of an explicit
finite-difference approximation. An explicit three-point smoothing for
dynamic variables was conducted for every time layer. The difference
scheme was constructed in the same manner as the known 'leapfrog'

350

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

scheme (see, e.g. Ref. 448). The computational errors were checked by
the mass and momentum integral disbalance, as well as by the comparison of results obtained, using networks with 80 and 200 nodes.
The parameters of BST potential were chosen as: n = 1.64, m = 3.1 and
b=4.91 x 10- 4 Additionally, the classic potential (n=2, b=O) was used.
The Mach number M was varied from 0.4 to 2, and for most results
the value r = 2.4 M was used. Parameter [3 was taken as either 0 or 1.
Figure 10.26 demonstrates the calculations with classic elastic potential, for time variations of a bar's free end, with several values of Mach
number. The increase in amplitude of oscillations is seen with increasing
values of Mo. Though in the initial compression phase of oscillation, the
effect of [3 is negligible, its infuence on the subsequent phases of oscillations becomes more important, and results in the augmentation of the
amplitude and a reduction in the oscillation frequency with growing [3.
This is explained by an increase in the relaxation time with an increase
in [3.
The effect of adhesion and separation conditions (10.98) on the bar's
dynamics is presented in Fig. 10.27. The relationships between the
dimensionless stress at the wall (jO(t) (solid lines), and the mass averaged
velocity U(t) (dashed lines), are shown here for f3 = 1 and various values

to
0.41--j'+-___,_----,----.J->O'cf---1,---~,____----___,_---___,

.--

--1,0

-0.2r-------'------\r'----~!--

Fig. 10.26. Time dependences of the dimensionless coordinate I(t) of the free
end of the bar. Curves 1, 2 and 3 correspond to M 0 = 0.4, 1.0 and 2.0; solid lines
for {3=0, dashed lines for {3= 1. 442

Flows Close to Simple Shear and Simple Extension

351

-10 '-----;-t.'-;;OSO.2-----;;Oc,..4--~O.6-

Fig. 10.27. Effect of the adhesion and separation conditions (eqn (10.98)). a O is
the dimensionless stress at the wall (---) and U is the mass average velocity
of the bar (- - - -). Curves 1, 2 and 3 correspond to M 0 = 0.4, 1.0 and 2.0,
respectively.442

of Mach number Mo. The shaded area corresponds to the criterion of


adhesion with the values k=0.35-3.5. When the value of wall stress (J(t)
exceeds a certain critical value 3ko from a given strip, which corresponds to the condition of separation (J(t*)=2k o , the bar is detached
from the wall with velocity U(t*). The more the Mach number, the
higher the velocity U(t*).
Distributions of the principal parameters, j, (J and v along the bar
length at different times are shown in Figs 10.28(a-c). The calculations
were carried out for classic elastic potential with values M 0= 1 and
fJ=1. The curves numbered 1-6 correspond to the times t 1 - t 6 . Time t1
is within the compression phase, t2 is the beginning of the compressed
bar unloading (where there is no shock wave but rather a diffuse front),
t3 and t4 represent the continuous unloading, and t5 and t6 represent
the stretching. The distributions of undisturbed values for 0 ~ x ~ 0.5 are
denoted by dashed lines. It is noteworthy that at time t1> all distributions are close to those computed for the shock wave propagating from

352

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

5
v

~-r----+--~

( c)

---- I

-01

-0.2

0.0

0.2

0.4

0.8

I
I

I
I
I

(b)
0.4

-0.2

6" -

I
I
I

0.2

Q4

:x:

2)

0.2

0.4

5-"" 4 - 3....-

/l
I

Fig. 10.28. Distributions of dimensionless variables along a bar: (a) crosssectional area f(x, t); (b) stress a(x, t); and (c) velocity v(x, t), for Mo= 1 and
fJ= 1. Curves 1-6 represent different dimensionless time instants tk =0.053,0.128,
0.205, 0.268, 0.310 and 0.465. 442

right to left. Also, the calculations with BST elastic potential made no great
difference to the results as compared to computations based on the classic
potential. The reason for that is that at the values of Mach number used,
the values of A in the extensional phase were not so high, but in the
compression phase both the potentials predict almost the same results.
More complicated behaviour was found for the case when preliminary
stretched viscoelastic bar strikes against a rigid obstacle. An interested
reader can find the results of these numerical simulations in Ref. 442.

353

Flows Close to Simple Shear and Simple Extension

10.10 THE RELATIONSHIP BETWEEN PURE SHEAR


(PLANAR EXTENSION) AND SIMPLE SHEAR
DEFORMATIONS FOR VISCOELASTIC LIQUIDS

Let a sheet of a material be uniformly stretched in the Xl direction,


freely contracted in the X2 direction orthogonal to its free surface and
be non-deformed in the X3 direction, orthogonal to Xl and X2 (Fig.
10.29). If, additionally, there is no volume variation, such a deformation
is called pure shear or planar extension (see e.g., Ref. 16, Section 5.3 and
Ref. 148, Section 10.6). In the case of continuous pure shear deformation with extension rate K(t), the total kinematic matrices have the
form:
C(t) = diag{C(t), C-l(t), I}

e(t)=K(t).diag{l, -1,0},

ro(t) = 0

(10.100)

where C(t) is the total Finger strain tensor in continuum.

Fig. 10.29. A schematic diagram of pure shear.

Now let us consider the pure shear flow of viscoelastic liquids. In the
case of constitutive eqns (1.71)-(1.76), all the elastic matrices Ck have a
diagonal form similar to that for the total Finger tensor C, so these
equations take the form: 38
O"u = ~)ftk/nk)(c~k-Cknk),
k

0"22=0,

W k = (ftk/2n~)( c~~ + Ck nk - 2)
0"33= L(ftk/nk)(l- ck nk)
k

(10.101)

It was taken into account that on the free surface of the sheet, the stress
component 0"22 =0.

354

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

For the steady isothermal pure shear flow with a constant extension
rate K, the first (evolution) equation in eqns (10.101) yields:
(10.102)
Then for small and large values of strain rate K, there are the following
asymptotic expressions for the stress components:

r 1 ~ 1 (1O.103a)
k
(J 11

L (.Uk/nk)Ck\

r N'P 1

(1O.103b)

The predictions of various viscoelastic constitutive models in pure shear


flow were recently discussed by Larson (see pp. 86 and 164 in Ref. 6).
There are not so many data available in the literature for pure shear
viscoelastic flow. Some preliminary experiments for the steady flow of PIB
were reported comparatively recently.449 It was shown in Ref. 449 that
(J11 and (J33 are increasing functions of K; the effective viscosity in
direction 1, (Jl1/K, being approximately constant and in direction 3, (J33/K
intensively decreasing. The formulae (10.103) can correspond to the data if
nk are equal or slightly more than unity, which is also in the agreement
with the simple extension data for polyisobutylene discussed in Chapter 6.
Now let us consider the difference between the pure and simple shear
deformations for the viscoelastic constitutive eqns (1.71)-(1.76) related to
parallel connected nonlinear Maxwellian modes. It should be recalled
that for elastic solids and Newtonian liquids, the simple and pure shear
coincide within a turn of the coordinate axis. As can be seen from eqns
(5.3)-(5.5), the set of equations for each kth Maxwellian mode for the
simple shear, can be reduced by orthogonal transformation
cos CfJk

- sin CfJk

qk = sin CfJk

cos CfJk

to the form (10.101) with K=Kk=(y/2) sinaCfJk. Here Ck is the maximum


principle value of elastic strain and CfJk is the rotation angle of coordinate
axes, whose time evolution is described as follows:
C; 1 dCk/dt + (c~ -1 )/(28~ Ck) =

Y sin 2CfJk
(10.104)

Flows Close to Simple Shear and Simple Extension

355

We now show that <fJk depends on the index k. For this purpose,
consider the steady isothermal simple shear with simplifying assumption
(J~=(Jk(T). In this case, eqns (10.104) are reduced to:

(10.105)
Formulae (10.105) show that the angle <fJk is determined by the dimensionless strain rate r k = (Jk y. This means that there is no unified
orthogonal transformation, independent of the number of relaxation
mode, which relates the simple and pure shears by a single rotation of
coordinate axes. The same results hold for the BKZ viscoelastic constitutive equations. 6 It should be noted that in the linear region of viscoelastic
deformations, where ck ....... 1, <fJk ....... n/4, as in the case of Newtonian liquid,
and such transformation is possible.

CHAPTER 11

Melt Flow Instabilities

In this chapter, we consider flows of polymers in dies and capillaries at


high Deborah numbers. The following sections present literature surveys
of experimental data and some of the results of theoretical studies of melt
flow instabilities, which are the most important effects in high Deborah
flows.
11.1 EXPERIMENTAL DATA

11.1.1 Historical
In 1942, Garvey et al. 450 were probably the first to describe and
photograph severe distortions, while extruding synthetic rubber compounds from dies. In thermoplastics, the first observation of extrudate
distortion is attributed to Nason 451 who, in 1945, extruded cellulose
acetate, PS and PV through a flat entry die with a small length to
diameter ratio. At a low shear rate, the extrudates were smooth; but at
higher rates they become rough and wavy. As the pressure drop was
increased, the roughness and waviness also increased. Nason noticed that
the waviness occurred at Reynolds numbers in the range 800-1000. Three
to five years later, Spencer and Dillon452-454 reported the flow of molten
PS through capillary dies with the aspect ratios ranging from 20 to 40.
Instability of flow was discovered at a constant value of the wall shear

Melt Flow Instabilities

357

stress of approximately 0.1 MPa. This critical shear stress was independent of temperature and inversely proportional to molecular weight. At
low shear stresses, the extrudate surface was smooth and the usual die
swell occurred. At higher shear stresses, Spencer and Dillon observed a
regular extrudate spiralling with some sort of surface roughness. At much
higher stresses, the surface roughness disappeared but the extrudate was
grossly distorted. Tordella,455 in 1956, gave the name 'melt fracture' to
the phenomenon. He observed various instabilities for PTFE, PE,
PMMA and Nylon 66. 344 ,456-459 In 1957, Clegg460 reported another
type of irregularity with wavelengths much smaller than the capillary
radius. This was later known as 'sharkskin'. Since then, many researchers
have observed surface and core irregularities of varying degrees in a
number of different polymers and have introduced a variety of names to
describe the phenomenon. Such names include 'elastic turbulence', 'waviness', 'ripple', 'bamboo effect', 'sausage link', 'matte', 'spurting', and lately,
'ring', 'crack' and 'cork flow'.
Several excellent reviews on the subject have been published; the most
recent being presented in Refs 461-463. Thus the present review is
focused on the basic effects and the recent results.
11.1.2 Extrudate Appearance
Many reports include photographs showing the appearance of the
irregularities. Some of them are shown in Fig. 11.1. 456 The type of
roughness varies considerably; the degree and extent of the distortion
depending on melt temperature, capillary geometry and polymer structure as well. Whereas the surface was usually smooth and glossy at a low
shear rate, the appearance at high shear rates was unpredictable. Some
extrudates were smooth and glossy at shear rates well above those where
melt fracture normally occurs. This was later attributed to 'superextrusion'. Bagley464 demonstrated that exterior surfaces could be smooth
although the interior showed evidence of melt fracture. Sometimes an
uneven or spiral appearance was readily visible, while in some other
cases, an apparently smooth surface would show imperfections when
viewed under low power magnification. Since there is no distinct criterion
to define surface roughness, it is impossible to find precisely the shear rate
at which it is initiated. Generally, the various irregularities can be
grouped into three classes:

Matte: a dull surface without gloss. It is considered that matte is


not initiated at a critical stress.

358

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Shear stress
(D cm~2)

2-6)( 1Q6

Fig. 11.1. Specimens of PMMA extrudate at various wall shear stresses. Conditions: 0.060 cm radius; 0.020 cm long capillary; T= 170C. The first irregularity
occurred at 4.8 x 10 6 dn cm 2 stress. The bottom specimen was broken in
handling. 456

Ripple: characterized by a series of peaks and valleys on the


extrudate surface. Generally, the magnitude of the roughness is
about 1/10 to 1/5 that of the overall specimen diameter. Sometimes, the surface form may resemble the thread of a machine screw
and, at times, sharkskin. There seems to be an almost common
tendency for the materials to form some type of spiral.
Gross (wavy and helical) irregularities: a most severe form of
distortion, wherein the wavelength of the distortion is commensurate with the diameter of the specimen. Commercial polymers
exhibiting wavy type of distortion include low-density polyethylene
(LDPE) (both long and short chain branching), polymethyl metacrylate (PMMA), polytetrafluoroethylene (PTFE), polyvinyl
butyral (PVB), polystyrene (PS), polypropylene (PP), polyvinyl
chloride (PVC) and various elastomers. In branched PE, (before

Melt Flow Instabilities

359

the surface of the extrudate shows any distortion, the core assumes
the form of a regular helix. 345 ,465 Linear PE shows a characteristic
screw-thread appearance. 461 Elastomers exhibit a quasi-thread
reminiscent of a cone within a cone. 461 Polyvinylidene fluoride
produces a double entwined helix. 459 Copolymers exhibiting irregularities include ethylene-propylene copolymer,461 and tetrafluoroethylene-hexafluoropropylene copolymer,45o both of which
exhibit screw-thread appearances. Styrene-butadiene copolymer is
grossly distorted over all ranges of shear rate even though PS 453
and PB 350 are known to display very regular extrudate distortions.
Further descriptions and details of a wide variety of polymer melts
can be found in the reviews by Tordella,459 White,461 Ballenger et
al. 350 and Bialas and White. 466 ,467
11.1.3 Superextrusion: Second Stable Regime
In 1960, Schreiber et al. 468 observed that at shear rates higher than those
resulting in gross extrudate distortion, a second stable flow regime may
occur where the extrudate would again be smooth. A 1961 DuPont
patent (VS Patent No.2 991 508 VI) indicated that for linear PE, there
is a shear rate range above the normal critical rate, where the extrudate
again becomes smooth, thus allowing a higher processing rate. The
phenomenon was reportedly utilized in high-speed blow moulding at
10000 s -1 .469 Tordella 458 ,459 has observed similar behaviour for linear
PE and tetrafluoroethylene-hexafluoropropylene copolymer. Although
the critical shear rate for PTFE is normally as low as 10 - 5 S - 1 at 380C,
at much higher shear rates smooth extrudates were obtained, generally at
high pressures combined with high temperatures. The explanation given
is that the melting point for PTFE is very pressure-sensitive, the gradient
being about 10 C for a change in pressure of 6.9 MPa compared with
2-3 C for many polymers. The effect occurs under conditions where the
high pressures increase the melting point above the processing temperature. As the PTFE flows down the capillary, it is decomposed and a
smooth extrudate is formed. The best results appear to be obtained at
high pressures (e.g. 69 MPa) and temperatures (380C).
Dennison469 stated that polymers such as branched PE, PP and PS do
not appear to show a second stable regime; however, Schreiber et al. 468
indicated that it should be possible for any polymer system to exhibit this
behaviour.
Tordella459 and Ramamurthy470 reported that a group of Japanese
researchers found a second region of stable flow for PE followed by a

360

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

second unstable region where the extrudate was distorted again. These
investigators claimed that linear PE requires low temperatures, large
capillary diameter and high shear rates to attain the second region of
stability. Yung 471 also observed a second region of stability in linear PE
and found that lower values of melt index seemed to favour the attainment of the second stability regime.
This process is particularly advantageous for such polymers as FEP,
whose onset in melt fracture happens at relatively low shear rates. In
superextrusion, the polymer melt is believed to slip relatively uniformly
along the die wall.
11.1.4 Change of Slope in Flow Curve
Apart from visual observation, the onset of melt flow instability has been
associated with a distinct change in the slope of the effective flow curve,
wall shear stress versus apparent shear rate. 456 Some researchers have
established that sharkskin surface melt fracture is accompanied by a
change in slope followed by a sharp increase at the onset of gross melt
fracture. 472 Others recognized that surface irregularities may not always
be detectable on the flow curve for every polymer susceptible to extrudate
irregularities, and thus suggested that a change in the slope of the flow
curve must occur around or at the onset of melt fracture, whether they
are the surface irregularities or gross distortions. 473 This change of slope
is real and is consistently reproducible for the many polymers investigated.
An examination of the published flow curves for such widely studied
polymers as both linear and branched PEs reveals, in many cases, a
change of slope around a value of wall shear stress of approximately
0.1-0.14 MPa. Above and below this value, a power law could be used to
accurately represent the steady flow data. For a narrow MWD HDPE at
190 0 e with a melt index (MI) equal to one, Lupton and Regester 474
reported a change in slope from 0.59 to 0.4 at wall shear stress of
0.14 MPa. At 160 o e, Metzner et al. 475 observed a change in slope from
0.49 to 0.32 for HP-LDPE at a wall stress approximately equal to
0.12 MPa. Also, in the tabulated corrected data of Bagley464 for both
HDPE and HP-LDPE, slope changes were revealed at a wall shear stress
of about 0.1 MPa. However, it is not clear from the published data
whether or not these investigators observed any surface irregularities at
stresses corresponding to the slope change.
On the other hand, Kurtz 476 indicated a change of slope at the point
where surface melt fracture is observed for three different LLDPEs with

Melt Flow Instabilities

361

a melt index of 0.5 to 2.0 and weight average MW ~0.93 x 10 5 -1.39 X 105
The measurement of onset and depth of melt fracture was done by optical
microscope. Kurtz determined the onset of melt fracture by the first sign
of a periodic surface wave under 40X magnification and side lighting.
This level of melt fracture was well below that seen with the naked eye or
even by surface roughness measurements.
Before any conclusion is reached, it should be pointed out that not all
polymers have a sharkskin surface during extrusion from dies. Sharkskin
melt fracture as described by Brydson477 and Cogswell,478 is generally
found with linear polymers of narrow MWD and in certain filled
polymers. In particular, unmodified linear low density polyethylene
(LLDPE) extruded through standard film dies displays sharkskin melt
fracture whereas low density polyethylene (LDPE) with long chain
branching does not.
Herranen and Savolaninen 479 utilized an ultrasonic method to observe
the melt fracture phenomenon on-line in the extrusion of LLDPE and its
blends with (branched) LDPE. The experiments were performed using a
Nokia MP40-24D extruder equipped with a special head for measuring
the melt temperature and pressure and to enable the ultrasonic velocity
in the melt to be determined. The Maddock-UC polyethylene and an
HDPE screw with two mixing sections were used in the experiments. The
melt temperatures were 237 and 267C. Unifos LLDPE 8020 (MI-0.9,
0.919) and Pekema LDPE B1222-05 (MI-1.2, 0.922) were the raw materials. The sharkskin surface appeared at a shear stress of 0.35 MPa. The
value of shear stress was independent of the screw geometry. At a melt
temperature of 267C, the critical shear rate was about 1000 s - 1 and it
decreased with decreasing temperature. Increasing the shear rate produced alternate wavy and smooth sections. At still higher shear rates this
behaviour ceased and the surface of the extrudate became completely
smooth, due to supposedly continuous slip along the surface of the die.
Above this super-flow region the second melt fracture occurred, with a
corresponding shear stress of 0.5 MPa, independent of temperature and
screw geometry. The critical shear rate was about 5000 s - 1 at a temperature of 267C and decreased to 3600 s -1 when the temperature was
decreased to 237C (Fig. 11.2).
Blending LDPE with LLDPE (10:90) reduced the shear rate and shear
stress at which the fine scale surface roughness was observed (Fig. 11.3).
The smooth super-flow region was narrower than that for pure LLDPE.
Melt fracture, however, appeared at the same shear rate for both pure
LLDPE and LLDPE/LDPE blends (90: 10). Increasing the amount of

362

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Fig. 11.2. Specimens of LLDPE extrudate at various shear rates at 237c. 479

LDPE in the blend caused the sharkskin to disappear and only gross
melt fracture was obtained.
Melt fracture of pure LDPE was observed at a shear rate of 1400 S-1
and shear stress of 0.21 MPa. The low value of the shear stress for LDPE
was as a result of the short length of the die. It was also demonstrated
that the occurrence of melt fracture can be correlated with the shape of
the shear rate versus shear stress curves.
Herranen and Savolaninen 479 reported that in asymmetric flow, the
extrudate distortion increases for branched LDPE as the die length is
increased. Increasing the length/distortion (L/D) ratio from 15 to 26 results
in an LDPE extrudate without any melt fracture while the melt distortion
of LLDPE moves to lower shear rates. Thus we can conclude that for
polymers that exhibit sharkskin, the onset of the instability is detectable by
a change of slope in the flow curve, and for polymers that do not exhibit

Melt Flow Instabilities

2 (a)

363

2 (b)

Shear stress ( kPa )

aJ

iiil_

10

21

.c

V1

Shear stress (kPa)

Fig. 11.3. Shear rate versus shear stress for ((a)-(d)) LLDPE and ((e)-(i))
LLDPE/LDPE blends--+sharkskin<-melt fracture. (a) HDPE screw at 237C; (b)
Maddock UC screw at 237C; (c) HDPE screw at 267C; (d) Maddock UC screw
at 267C. All data for (e)-(i) were obtained for HDPE screw at 267"C with an
LLDPE/LDPE content ratio equal to: (e) 0.9:0.1; (f) 0.75:0.25; (g) 0.5:0.5; (h)
0.25 :0.75, (i) 1.0:0.0. 479

this, a change in the slope of the flow curve indicates the onset of gross melt
fracture. It is known that even for polymers that exhibit sharkskin, the
temperature and upstream flow conditions are crucial. Thus a critical shear
stress, weakly affected by temperature and molecular weight, will trigger off
sharkskin and it will be noticed as a change of slope in the flow curve.

364

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

11.1.5 Hysteresis
In 1958, Bagley et al. 480 first reported the occurrence of a doubly
branched flow curve for linear HDPE extruded through a capillary at
constant pressure. At the critical pressure drop, the output became
double valued and a hysteresis effect was noticed. When the pressure
drop increases, a point is reached at which the output jumps from a low
to a high value. With further increases in pressure drop, the output
continued to increase at the higher level. As the pressure was decreased,
the output decreased until a point where it jumped to a lower value.
The pressure at which this jump occurred was lower than the pressure
at which the jump to the higher output occurred. Tordella345.455 observed this hysteresis effect for both linear polyethylene and tetrafluoroethylene-hexafluoropropylene copolymer. Lupton and Regester 474
also observed this phenomenon while extruding HDPE in a constant
rate capillary rheometer. They noted that under either constant stress or
constant rate operations, there are some parts of the flow curve (Fig.
11.4) that are related to a steady flow. However, steady flow cannot be
maintained on the left branch above an 'upper apparent critical shear
stress', Seu, or on the right branch below a 'lower apparent critical shear
stress', Sci' If either of these cases is attempted, the flow shifts to the
other branch. Kissi and Piau 472 also reported hysteresis in the flow
curves in both the pressure and flow-rate controlled capillary
rheometers, for a series of highly entangled PDMS. Uhland 481 has also
reported the same phenomenon. Myerholtz 482 was seemingly the first to
demonstrate the onset and increase in the hysteresis loop with the
increasing narrowness of MWD for HDPEs and their copolymers
(Fig. 11.5).

Fig. 11.4. Hysteresis in a technical flow curve for a capillary flow of PE at


190C.474

Melt Flow Instabilities

365

Shear r a t e -

Fig. 11.5.

Effect of MWD on hysteresis in a technical flow curve for the


capillary flow of HDPEs at 190C, L/D=40. 482

11.1.6 Oscillations
Tordella 456 pointed out that at the onset of melt flow instability, the
volumetric flow rate becomes irregular. He did not indicate, however,
whether the experiment was conducted at a constant pressure or at a
constant flow rate, and it is assumed that a constant flow rate was used.
Metzger et al. 483 studying the flow behaviour of low molecular weight
PE, showed the plots of load in a constant-rate capillary rheometer as the
shear rate was increased and decreased. The loads were constant in value
whether the melt was fracturing or not. Thus, it was not possible to detect
melt fracture from the loads recorded. The polymer exhibited melt
fracture but there were no inflections or discontinuities in the curves, even
when the apparent viscosity was increased by reducing the melt temperature. On the other hand, when high molecular weight HDPEs were
extruded through capillaries, it was noted 480 that at high shear rates, the
load no longer remained constant but oscillated between two values; both
the upper and lower load values remaining essentially constant. If the
piston's velocity was increased further, beyond that at which the load
oscillates, then steady conditions occurred again.
Metzger et al. 483 pointed out that under conditions of oscillating loads
(Fig. 11.6), the appearance of the extrudate alternated between smooth
and rough, or it was continuously rough with the degree of roughness
varying. In some cases, the extrudate remained smooth, but based on the
high and low readings, alternate stress appeared to be double-valued.
Since load oscillation may occur at several shear rates, the flow curve
would then contain an envelope. Furthermore when the load cycled, the

366

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

I1J

.9

Time_

Fig. 11.6. Typical oscillations of pressure in a pre-capillary reservoir. 483

extrudate pulsated, i.e. although the piston travelled at a constant speed,


the shear rate was not constant. Under oscillation of the load, the
appearance of the extrudate changed markedly and at a constant piston
velocity, the extrudate consisted of uniform but alternating sections. It is
pertinent to note that under conditions of load oscillation, the shear rate
oscillated between two distinct values. The flow curve was discontinuous
and there was a range where the melt cycled between two average shear
rates. An attempt to operate within this region resulted in a non-cyclic
oscillating flow. Outside the unstable range, on either side, the flow rate
was constant. It was noted that melt fracture occurred at shear rates well
below those where the melt flow fluctuated. This observation led to the
conclusion that melt fracture and the oscillations are probably separate
phenomena or at least separable degrees of the same effect. The pulsatile
flow appears to be associated with a slip effect and with melt compressibility; however, its relationship with the onset of melt flow instability is
not so clear. Those instabilities can occur either below or above the
'forbidden' range although it may well be triggered by similar causes to
those which initiate pulsatile flow. 480
Oscillating flow has also been reported by Lupton and Regester. 474
Figure 11. 7(a) shows that pressure in their rate-controlled capillary
instrument oscillated, instead of changing directly to the normal steady
pressure indicated by the dashed line in this figure. In oscillating flows,
the extrudate was usually smooth through all of, or all but the last part
of, the ascending sections of the pressure-time curves and rough through
all of or all but the last part of the descending sections. In extruders,
oscillating flow can usually be detected by visible shifts in the rate at

367

Melt Flow Instabilities

'iii

.9- 4000
~

:J

Oscillating flow

d'=2200

1
L

2000

0::

/~

S-1

Steady flow

Time (min)
(a)

'iii

.9~ 4000

a.

a;

t2000
E

0::

.. /4-1 min --.j

;1/------

Time
(b)
'~
~4oo0

:J

<11

a; 2000

0::

0.1

Time (min)

0.2

(c)

Fig. 11.7. Oscillating capillary flow of PE at 190C, LID = 16. (a), (b) and (c)
correspond to various apparent flow rates Q in the hysteresis loop shown in
Fig. 11.4.474 (1 psi = 6895 Pa.)

which extrudate emerges from the die and recurring changes in the diameter
and surface characteristics of the extrudate, and through oscillations in
pressure head. Other investigators have also noticed this oscillation

368

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

phenomenon. Kalika and Denn 484 observed that the onset of stick-slip
melt fracture is also accompanied by periodic fluctuations in pressure when
the rheometer is operated in the constant plunger velocity mode. Uhland,481 Lim and Schowalter485 and Kissi and Piau 472 have also mentioned that for a controlled average rate of flow, the instability is
accompanied by oscillations in instantaneous pressure and flow rate.
Ballenger et al. 347 den Otter 340 and Bagley and Schreiber486.487 all have
observed oscillations of stress at the critical rates with constant rate instruments. Den Otter indicated that LDPE did not exhibit oscillations of stress
at the critical stress. It seems that polymers that exhibit branched polyethylene-type behaviour do not appear to exhibit discontinuity in their flow
curves, but exhibit pressure fluctuations instead. Ballenger and White 336
reported the pressure fluctuations in a constant-rate experiment with PS.
Typically, the period of oscillation is of the order of 1 s. In 1967,
Myerholtz 482 showed that the period of oscillation increases with as the
plunger speed increases.
There are also other aspects of the oscillating flow phenomena, which
can be summarized as follows.
(1) The effects seem most marked with HDPE. There is, however,

(2)

(3)
(4)

(5)

some indication of a similar effect with PTFE, narrow molecular


weight distribution PP, and PB.
Cogswe1l 488 observed almost the same effects at similar critical
stresses in flow between concentric cylinders.
He also showed 488 that the effect is not seen with zero land-length
dies.
With HDPE, the critical shear rate for the onset of oscillatory flow
decreases with increasing average molecular weight. Broadening of
the molecular weight distribution also markedly increases this
critical shear rate.
The upper critical shear rate at which pulsing flow disappears was
observed by Myerholtz 482 as seemingly independent of the molecular parameters.

11.1.7 Site for Initiation of Instability


Until now there has been no agreement even on the site or sites of
initiation of either type of instability. There are only three possible
regions of initiation of flow instability.
Entry region: This region begins where the flow first starts to
accelerate prior to entering the contraction, and ends when fully

Melt Flow Instabilities

369

developed laminar flow is established downstream of the contraction.


Land region: This region contains the fully developed flow region
between the die entry and exit.
Exit region: This region begins where the fully developed flow is
lost prior to leaving the die.
Flow in both entry and exit regions contains an extensional or
elongational component due to the acceleration/deceleration of flow in
the respective regions. The land region is essentially characterized by
shear flow.
Upstream of axisymmetric dies, the material undergoes acceleration as
it passes from a large duct cross-section to a smaller one. Here, along the
walls, there is shear flow and along the axis it is elongational. Thus the
entry flow is of the mixed type, and both elongational and shear
viscometric functions are expected to be important. Also, the flow here is
unsteady from the Lagrangian viewpoint and start-up properties seem to
be important, too. This means that while transient, shearing and elongational properties cannot always be easily measured over the whole range
of interest, they should be closely examined when choosing the constitutive equation, especially in modelling and simulation.
11.1.8 Structure of Polymer
Significant differences have been noticed in the unstable behaviour of
linear and branched polymers. In particular, the following differences
have been noticed between HDPE and LDPE
(1) large recirculating regions in the entry region for LDPE but not for

HDPE;489
(2) increasing the die length has resulted in a decrease in severity of
distortion for LDPE but the reverse for HDPE;
(3) discontinuity in the flow curve for HDPE but not for LDPE.

These observations tend to suggest that the onset of extrudate distortion for LDPE is related to disturbances in the entry region 489 while for
HDPE it has often been suggested that a stick-slip mechanism at the die
land causes the instability.490,491
It should be noted, however, that there are a lot of confusing results for
other polymers, as related to their linear and branched structure, which
are not so clear as in the case of polyethylenes (see, e.g. the review in Ref.
462). Also, there are a lot of indications that polymer MWD plays a very

370

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

important role in melt flow instabilities. As demonstrated in Fig. 11.5 by


Myerholtz,482 only the polymers with narrow MWDs exhibit the hysteresis and associated with that, spurt flow instabilities and oscillations.
These phenomena were not observed for polymers with broad MWDs.

11.2 SHARKSKIN AND SPURT INSTABILITIES

Melt fracture is a generic name used to describe different irregularities


and instabilities which produce distortions or non-smooth surfaces in
extrudates. Many researches distinguish between sharkskin melt fracture
and spurt or stick-slip melt fracture,462.4 76-478 but some others prefer to
lump all the flow instabilities together as one phenomenon under the
term 'melt fracture', In this subsection, an attempt is made to separate
these types of melt flow instabilities.
11.2.1 Sharkskin

Sharkskin is characterized by: (i) steady flow through a die, (ii) a repeated
pattern of a wavy or ridged surface perpendicular to the flow direction on
the polymer extrudate, and (iii) a central core that is uniform in its stress
distribution in the axial or flow direction. In its milder form, it is known
as mattness or matte where the extrudate has a lack of surface gloss.
Rubber extrudates are invariably matte in surface appearance.
Sharkskin has been distinguished by Howells and Benbow 492 from
gross melt fracture, for the following reasons.
(1) Sharkskin has a perpendicular distortion whereas the gross disturbances usually have helical or irregular patterns.
(2) Sharkskin can occur at lower extrusion rates than gross melt
fracture and appears to be a function of the linear output rate
Q/nR2 as opposed to the critical apparent wall shear rate 4Q/nR 3
for the gross distortions.
(3) Sharkskin is very temperature dependent, being considerably delayed by an increase in temperature.492.493 Cogswe1l494.495 observed that die tip cooling can be used to eliminate sharkskin for a
wide range of polymers.
(4) Lubricating the die or use of PTFE in the die will lessen or
eliminate sharkskin. 476
(5) Sharkskin appears to be unaffected by die entry angle, the L/D
ratio of the die, and the material of construction of the die. A slight

Melt Flow Instabilities

371

improvement with a shorter die land was also observed.496.497


Kurtz 476 reported that the change in the level of sharkskin for
convergent or divergent angles up to about 10 at the die exit is
negligible.
(6) Sharkskin is affected by the die gap.476
(7) Molecular weight seemingly has little influence on sharkskin
whereas the MWD appears to be all important. Both by comparing fractions with both whole polymers 492 and by artificially
increasing the MWD by mixing polymers of widely varying molecular weight,479,498 it has been found that the narrower the
MWD, the greater the sensibility to sharkskin.
(8) The higher the swelling ratio and the lower the value of the end
correction factor, the less the degree of sharkskin.
Sharkskin is generally found in linear polymers of narrow MWD and
in certain filled or modified polymers. For example, modified LLDPE
extruded through standard film dies does display sharkskin whereas
LDPE with long chain branching does not.
Modification of the polymer structure to avoid sharkskin can be
affected by (a) changing the molecular weight, and (b) the MWD or the
degree of branching. These methods, however, tend to reduce the physical
properties.
Pearson 499 reviewed a number of the processes leading to irregular
surfaces in extrusion through dies and concluded that exit effects do have
separate and distinct results. For example, the dependence on extrudate
velocity indicates a surface effect. The dependence on temperature and
the relationship to die swell indicate the elastic effects, or the importance
of balance between elastic and viscous properties.
The mechanism of sharkskin has been variously associated with (a)
flow upstream of the die exit, (b) acceleration (stretching) at the exit, and
also (c) the abrupt edge geometry at the exit. The source of this effect is
thought to be a stress relief at the exit border line where the melt parts
contact with the die, usually the die exit. Kurtz 476 suggested that the
theory of sharkskin must encompass the rheological properties of polymers in shear as well as in extension. He proposed the following
two-point criteria for sharkskin:
(1) A critical wall shear stress defined by the point where the change of
slope of the flow curve occurs must be exceeded.
(2) An excess of a critical acceleration and stretching of the extrudate
surface for a sufficient time period.

372

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The second criterion may often be approximated by a critical average


velocity.
Taking advantage of the fact that, LLDPE shows no strain hardening
compared to LDPE, Kurtz,476 using the above criteria, was able to
establish sharkskin limits for various output rates, melt temperatures and
resin melt indexes.
11.2.2 Spurt Melt Fracture
In contrast to sharkskin, the more severe phenomenon of spurt melt
fracture shows a marked fluctuation in head pressure and a non-uniform
core axial stress.
The first fundamental results in studying spurt phenomena were
obtained in works by Lupton and Regester474 and Myerholtz. 482 These
results were further extended by Vinogradov and co_workers,183,281,291293,500-509 who worked with very narrow distributed polybutadiens (PB)
and polyisoprens (PI). Typical 'spurt flow curves' for a PB, shown in Fig.
11.8,508 show the hysteresis region,

10-8

(//j/
,
I

"

I:
"

10

20

30

40

P(MPa)

Fig. 11.S. Technical flow curves for spurt flow of polybutadiene PB-3 at 20 e in
capillaries having a diameter D = 1 mm and lengths L of (1) 1.2, (2) 5.4, (3) 9.9,
(4) 15.0 and (5) 20.4 mm.
0

Spurt phenomena can be characterized in part by alternating high and


low mass flow rates with pressure drop following the same periodicity.472,474 High-frequency pressure fluctuations with a central core of the
extrudate being generally inhomogeneous in all directions, are the characteristic features of the spurt phenomena. Unlike sharkskin, the extrudate, though grossly distorted, exhibits a glossy surface. Both spurt

373

Melt Flow Instabilities

fracture and elastic turbulence do depend on die entry angles, and the
LID of the die.
Hatzikiriakos and Dealy recently reported the occurrence of
slip and melt fracture for HDPE using capillary and slit rheometers 510
and a new developed sliding plate rheometer. 511 The HDPE used had
MW=1.79 x 105 and MW/MN=9.4. They observed a critical shear
stress of 0.09 MPa for the onset of slip and melt fracture in the three
instruments. They carried out steady and oscillatory shear experiments
using a Rheometries Mechanical Spectrometer and discovered that not
only the critical shear stress is independent of frequency, but its value still
remains approximately the same, equal to 0.09 MPa, as obtained for
steady shear experiments. Their results depicted in Figs 11.9 and 11.10
demonstrate that the presence of fluoroelastomer coating the interface in
the slit and capillary dies makes about 80% reduction of the critical stress
for the onset of slip, and also altered the relationship between the slip
velocity and wall shear stress. According to their report, the critical shear
stress for the onset of extrudate distortion was unaffected. The results
implied that while slip may be occasioned or accompanied by the
breakdown of adhesion at the polymer wall interface, extrudate distortion
is not primarily due to the occurrence of melt slip. They thus distinguished between the critical shear stress for the onset of slip and that for
the onset of melt fracture. While the critical shear stress for the onset of
slip is independent of temperature, that for the onset of melt fracture is a
0.10 T=180 0 C,h=0.36mm
Resin 569/3830

O.OB

I:.

'& 0.06
2

0'+J

O~

o~

0.04

0.02

l"

Jeri'

I:.

01:.

o
o

ceo

0000

4:P 0 Plates untreated


I:. Moving plate treated (DFL spray)
o 90th plates treated (DFL spray)
o Stationary plate treated
( Dynamar 9613)

0~~10~~2~0~~3~0~-4~0~~5~0~~60
Yn(s-1 )

Fig. 11.9. The effect of fluorocarbon (Dynamar 9613) on the adhesion of HDPE
56Bj3830.510

374

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Plate coated with Dynamar 9613


T=180C
'"' 0.010

[J
[J

0.001 '--_ _ _-'-_ _-'-_--::-~-,........


0.03
Shear stress (MPa)
Fig. 11.10. Slip velocity as a function of wall shear stresses for a surface treated
with Dynamar 9613 at 180c. 510

function of temperature, the diameter of the die and the LID ratio. They
also reported that the action of the fluoroelastomer was very different
from that of a lubricant layer. The lubricant layer was found to shift the
flow curve downwards at all shear rates used when such a critical shear
stress for the onset of slip did not exist (Fig. 11.10).
The results from the sliding plate rheometer are, however, not as
definite. One fluorocarbon (DFL Spray) reduces the slip velocity but does
not affect the critical shear stress for the onset of slip, while another
fluorocarbon (Dynamar 9613) increases the slip velocity greatly and
reduces dramatically the critical shear stress for the onset of slip, from
0.09 MPa to 0.025 MPa.
These findings are very important and, if independently confirmed,
could resolve in perspective a number of contradictions on the melt
fracture and the stick-slip. Lowering the critical stress for the onset of slip
is attributed to the fact that the presence of a layer of fluorocarbon alters
the adhesion characteristics of the interface. The extent of the onset of
slip, for both the capillary and slit rheometers, from 0.109 MPa to
0.02 MPa and from 0.09 MPa to 0.025 MPa, respectively, is significant.
To the authors' knowledge, no other study, except that of Vinogradov et
al. 508 has yet presented such a result, giving a quantitative measure of the
effect of fluoroelastomer on the critical rates.

375

Melt Flow Instabilities

Vinogradov et al. 508 also presented flow curves (shown in Fig. 11.11) for
well characterized polybutadiene PB-3 flowing through steel, glass and
Teflon capillaries. At low shear rates, the flow curves describe Newtonian
behaviour, coinciding for all types of capillaries. The steel capillaries began
to show the deviations from Newtonian flow roughly at critical wall shear
stress 'spo However, in Teflon capillaries, this deviation occurs at
'er ~ 'sp/5; the polymer shear rate increasing faster than the shear stress.
Also, spurting began to occur at shear stresses Lcr ~Lsp/5 at the channel
exit for Teflon capillaries. This is because the outer edge, being a stress
riser, creates large and prominent transverse grooves on the extrudate
during spurting. Since the adhesion of PB to Teflon is much poorer, the
spurting at the outlet edge propagates into the channel to a considerable
depth. The polymer flows at the channel inlet and slides over the wall near
the exit. It is further shown that the stress value calculated with due
account for the entire channel length turns out to be lower as compared to
the actual one, and an apparent viscosity anomaly may seem to occur.
When the stress approaches Ls l, sliding begins even near the channel inlet.
This is indicated by the fact that the flow rates through a Teflon capillary
exceed those during spurting in steel and glass capillaries. Once the shear
stress Lsl is reached, PB slides perceptibly over the Teflon capillary surface.

Fig. 11.11. Technical flow curves


for polybutadiene PB-3 at 20 e in
steel (e), glass (< and Teflon capillaries: monolithic Teflon capillary with
L/D=6 (0) and steel capillary with
Teflon coating, L/D=4 (6). The arrows show the values of the spurt
shear stress. 508
0

It is surprising that Hatzikiriakos and Dealy's data510.511 contradict


the conclusions of Ramamurthy470,473,512-514 and those of Kalika and
Denn.484 Ramamurthy and co-workers lumped together several categories of extrudate distortion and considered each irregularity as a part of
general behaviour associated with polymer wall slippage. They argued

376

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

that the onset of sharkskin, for example, occurs as a result of the failure of
adhesion. In particular, Ramamurthy made the following observations
and conclusions.
(1) The applicability of the no slip at the rigid boundary assumption is
limited to a certain critical shear stress below which the extrudates
are smooth and glossy. The assumption is not valid when either
surface or gross irregularities are present.
(2) Loss of extrudate gloss and the initiation of slip in the die land
region, which occurs at a critical shear stress with an approximation value of 0.1-0.14 MPa, denotes the onset of extrudate
irregularities.
(3) Within the range of variables examined, the critical shear stress is
independent of the molecular structure (MW, MWD and branching), melt temperature, and the detailed design of the capillary.
(4) Failure of adhesion at the polymer/metal interface in the die land
region at the critical stress appears to be primarily responsible for
the initiation of slip and the observed extrudate irregularities.
(5) Under conditions of continuous blown film fabrication, some
methods providing good adhesion at the polymer/metal interface
by proper choice of materials of construction for the die land
region and/or the use of adhesion promoters in the resin, essentially eliminate the rate limiting effects of melt fracture.
(6) Standard capillary rheometer methods are unable to delineate the
influence of materials of construction and/or adhesion promoters on
melt fracture behaviour of polyethylenes. The effect of such variables
should be determined under actual fabrication conditions.
Kalika and Denn 484 virtually agreed with Ramamurthy's conclusions.
They pointed out:
'Our results are consistent with those of Ramamurthy: there is evidence
that the onset of sharkskin and the apparent curvature of the flow
curve are both caused by slip at the capillary surface. We observe more
intense slip with increasing throughput, and there is an apparent
periodic adhesive failure and reattachment on the scale of the residence
time in the capillary; multiple periodicities presented in the stick-slip
flow regime are associated with compressibility of the melt in the
reservoir.'
Kalika and Denn categorically stated that the sharkskin region also
demonstrates slip, but of lower intensity, and possibly continuously. They

Melt Flow Instabilities

377

admitted that the precision of their experiments does not allow them to
reveal any periodicities in the pressure curve for the region. They
surmised about the stick-slip melt fracture that a catastrophic failure of
adhesion at the melt/metal interface occurred, characterized by nearly
complete slip for a period of the order of a residence time in the capillary.
Despite the possibility of stress relaxation following failure, re-adhesion
does not occur until the stressed material near the wall has been removed
from the capillary. This may happen because of a free-energy change
affected by the mechanism of desorption (diffusion from the bulk being
too slow, so that convective flow is the only possible replacement
mechanism). The new material adheres (with a small amount of sharkskin-type slippage) until the stresses build up again to the point of
catastrophic failure.
Piau, Kissi and Tremblay 515,516 argued that the conclusions reached
by Ramamurthy and co-workers are not consistent with their experimental findings. They studied the flow of four PDMS through thin-walled
orifice dies (0.1 <L/D< 1), and clearly detected surface defects in the form
of scratches at the free surface which were capable of evolving into more
severe forms of extrudate distortions, called 'loss of gloss' or 'sharkskin'.
They also observed melt fracture for all the silicones studied and from
their visualization of the upstream flow concluded that the surface defect
even in its more severe form as sharkskin is an exit phenomenon related
to relaxation of stretching at the orifice outlet. This was because flow
upstream remains perfectly stable as long as the rupture regime is not
reached. On the other hand, instabilities were generated in the upstream
region for the flow regime at which the melt fracture appears downstream. Specifically, they argued:
(1) For very short capillaries and moderately high molecular weight
silicones with minimum entanglements, wall slippage is a minor
phenomenon, and cannot be responsible for the observed extrudate's defects.
(2) The onset of sharkskin in extrudate is not controlled by a critical
wall shear stress: the stress thresholds vary with the fluids and the
die geometry used and therefore cannot be considered as a criterion for the onset of surface defects.
(3) The onset of gross melt fracture defects appears to be controlled by
a stress critical value. This value varies according to the fluid
considered but is independent of the diameter and length of the die
for a given fluid.

378

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Piau et al. 516 further carried out extensive flow visualization studies of
slightly, moderately and highly entangled PDMS in capillary dies of
appreciable aspect ratios. Three well characterized silicone gums PDMS
were used (BG, LG2 and LG3). These had a critical molecular weight
MWc of the order of 27 000. This gives a rough estimation for the MW e ,
molecular weight between entanglements, two to three times lower than
the MW c' Thus Piau et al. satisfied the criterion given by Vinogradov et
al. 50B that for polydispersed polymers with MWjMN~ 1, slip occurs only
if the ratio MWjMW e > 10. Piau et al. carefully photographed the
extrudate distortions and their transition from one form to another,
covering scratch, crack, sharkskin, cork flow and chaotic melt
fracture, for slightly entangled (MW jMW c = 5), moderately entangled
(MWjMWc = 16), and highly entangled PDMS (MWjMWc >30).
They also extended the observations of Vinogradov et al. SOB to other,
particulary highly entangled, PDMS polymers and demonstrated the role
of test apparatus rather than that of construction material. Their conclusions include the following.
There is a difference in flow structure depending on the type of
capillary experiment. As shown first by Lupton and Regester,474 fluid
compressibility is a major parameter when screw or piston speed is
applied to the fluid in the flow rate controlled system, whereas it is
insignificant in the pressure drop controlled instrument. For pressurecontrolled conditions, slip appears simultaneously with a sudden increase
in the flow rate, and the flow curve exhibits a hysteresis regime. Under
flow-controlled conditions, slip is accompanied by oscillations of the flow
rate and pressure drop around their mean values, as a result of polymer
compressibility. Specifically,
(a) in the case of controlled pressure flow, a major jump in the flow
rate is seen on the curves while the cracks and sharkskin disappear
suddenly;
(b) in the case of speed-controlled flow there are oscillations in the
actual pressure between two flow phases: during the compression
phase, pressure increases with time and the actual flow rate is low,
the polymer adheres to the capillary wall and cracks as it leaves the
die; during the expansion phase, when the pressure drop decreases
with time and the actual flow rate is high, slip occurs and the
extrudate is virtually but not totally smooth.
The sizes of the sections disturbed by the sharkskin effect, corresponding to polymer adhesion on the wall, and of the smooth sections,

Melt Flow Instabilities

379

corresponding to polymer slip along the wall, depend on the length and
frequency of the compression and the expansion phases at a given
average flow rate.
In the cases of lower and very high flow rates, the experimental
apparatus appeared to have no effect. At low flow rates, the fluid adheres
to the wall and the sharkskin appearance is observed. At high flow rates,
slip may occur at the wall and this, in addition to the upstream
hydrodynamic instabilities, affects the appearance of the extrudate.
The following conditions for slip are established:
(a) the greater the MW and the smaller the LID ratio, the lower the
pressure at which slip occurs;
(b) the greater the MW, MWD, and LID ratio, the greater the jump in
flow rate generated by the occurrence of slip;
(c) for a given PDMS and die diameter, unstable flow corresponding
to the increase in flow rate or to the discontinuity in the flow curve,
is characterized by the mean mass flow rate independent of the die
length considered;
(d) for a given PDMS, the curve representing variations in mean flow
velocity V as a function of critical shear stress L Sp ' for the flow
regime with slip is independent of the dimensions of the capillary
(Fig. 11.12).
From the review of experimental effects in melt flow instability one
obtains the impression that the disturbances in capillary entry region are
the major factor for the polymers whose elongational viscosity increases
as a function of the extensional strain rate (e.g. LDPE). Other polymers
can be subdivided into two groups. For those that have relatively broad
MWD, only the sharkskin effect produced in the exit region is seemingly
possible (e.g. HDPE with wide MWD); and those with very narrow
MWD will display hysteresis, wall slip and oscillations in the land region
(e.g. narrow distributed polybutadienes).
11.3 MECHANISMS OF MELT FLOW INSTABILITIES
11.3.1 Viscoelastic Criterion for the Onset of Unstable Flow
Even in the early stages of the study of melt fracture it was concluded that
for polymer melts, inertia terms were insignificant. 456 Thus in 1963, the
characteristic number, r = eVil was introduced in Refs 17 and 18 (e, V
and 1 being characteristic relaxation time, velocity and length), which

380

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

102

Gum LG2:
Tungsten carbide
capillaries
L/O =2010.5
[J L/O =1010.5
A L/O = 20/2

...I

1d

til

E
E
~

10

1C51

10-1
t~ (bar)

10

Fig. 11.12. Average velocity V versus wall shear stress ri for the flow of LG2
through capillaries of different dimensions. 472

describes the elastic properties of polymeric liquids as a criterion of the


onset of melt flow instabilities. Reiner 19 introduced this parameter once
again in 1964 and termed it the Deborah number, De. Furthermore, the
same criterion was introduced in Refs 517 and 518. References 17, 18 and
517-519 showed that this criterion holds generally for all the polymers
with variations in values of De from 1 to 10. However, it is possible that
this criterion is not the only one to describe the onset of melt flow
instabilities.
11.3.2 Fluid Mechanical Instability
Any constitutive equation which claims to describe the viscoelastic
phenomena, has to possess suitable nonlinearities. Thus, as compared
to common hydrodynamic instabilities for Couette and Poiseuille flows
of Newtonian liquids, new kinds of fluid mechanical instabilities may,
in principle, occur in these viscoelastic flows with usual stick condi-

Melt Flow Instabilities

381

tions at the walls. However, the conditions of unstable flow will now
depend on the Deborah number but not (or weakly) on the Reynolds
number. A considerable number of theoretical investigations of the
problem reviewed in Ref. 462 have demonstrated the strong dependence of the instability on the specifications of the constitutive equations involved. Since at present there are many such equations, the
authors consider that further discussion of the problem here would be
futile.
It should also be mentioned that attempts are being made to relate the
instabilities observed in the experiments with the internal instabilities of
viscoelastic constitutive equations (for a review see, e.g. Ref. 520). As
shown in Chapter 3, all these instabilities have no fundamental reason
and are seemingly related to poor modelling of different terms in
constitutive equations.
11.3.3 Fracture (Discontinuity) of Polymer Melts in Flow
Hutton 521 ,522 proposed the mechanism of onset of melt flow instability
owing to crack and discontinuity formation in the polymer flows under
elevated stresses. Despite the criticism,459,523 this idea warrants consideration since, in practice, there are many cases in which polymer extrudate
breaks into a number of almost solid pieces. In Ref. 524 some suggestions
were made on how to improve Hutton's approach.
11.3.4 Wall Slip Effects
This is the most popular mechanism that might be the cause of many
types of melt flow instabilities. It was first suggested by Benbow and
co_workers 490 ,491,525 and broadly discussed in Section 11.2.1. This mechanism is related to the occurrence of hysteresis on the effective capillary
flow curve, to the spurt flow phenomena and oscillations of flow
parameters. Direct laser-Doppler measurements of flow slip in unstable
regimes of flow of PIB were first performed by Paskhin. 181 The theoretical studies (see, e.g. Ref. 462) suggest, however, that the wall slip produce
instabilities in flow only if there is a decreasing branch of wall shear stress
on the slip velocity. We will discuss in more detail the physical reasons for
the occurrence of wall slip phenomena, along with some theoretical
models of spurt flow effects in Section 11.5.
11.3.5 Other Explanations and Assumptions
At the end of the 1950s Spenser and Dillon452 .453 proposed a mechanism
for the buckling of polymer extrudate due to differences in the orientation

382

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

of polymer parts in the near-wall region and core of flow. Though this
cannot explain the occurrence of melt flow instability in the land region,
Vinogradov et al. 508 argued that the mechanism is able to explain the
flow events in the exit zone of capillary flow.
Tordella 459 also suggested viscous heat generation as a possible source
of disturbances. Though the studies of non-isothermal effects considered
in Chapter 9 do not deny such a possibility, Lupton and Register 474 and
Clegg 526 have presented convincing evidence that these effects, at least for
studied polyethylenes, are negligible.
Overdiep and Van Krevelen 314 suggested the onset of instability as a
result of a loss of concavity in the velocity profile in flowing polymers.
These effects have been considered in Chapter 8 when discussing the
Poiseuille flow with hardening phenomena. As shown by Paskhin,181
these effects usually occur on the background of sharkskin distortions of
extrudates.
11.3.6 A Concept of Polymer Fluidity Loss in Intense Flows
In Refs 291-294 a concept was formulated that under large strain rates
polymer liquids could lose their fluidity because of deformation orientation, and so gain the properties reminiscent of polymer behaviour in
the solid rubber-like or even glassy states. Such a 'transition' of trivial
form can be established in pre-steady situations in which the characteristic flow time llV is less than characteristic relaxation time
8. 517 ,518 Another and less trivial situation may arise in steady flows of
polymers and be induced by the increase in relaxation time due to
orientational phenomena. The physical ideas of this 'relaxation transition' have been discussed in relation to the hardening phenomena and
the possible methods of their mathematical modelling 8,65 in Section 7.4.
In high Deborah capillary flows of polymers, the loss of fluidity may be
displayed either in slip of polymer along the wall or (if the adhesion of
polymer to wall is high enough) in a cohesion rupture with crack
formation. On the contrary, in extensional flows of polymers, the loss of
fluidity will result in the fracture of the extended specimen. The ruptures of stretched liquid polymers, observed by Vinogradov and coworkers 5oo ,503,528 is discussed in Chapter 12.
The concept of fluidity loss, being not so developed theoretically, still
gives an opportunity for a uniform, at least qualitative, explanation of a
number of unusual effects observed in flows of polymers at high Deborah
numbers.

383

Melt Flow Instabilities

11.4 SLIDING FRICTION OF CROSSLINKED ELASTOMERS


AND WALL SLIP OF POLYMER MELTS
Prior to modelling spurt flow for polymer melts, it seems very useful to
discuss some results of experimental and theoretical studies developed in
sliding friction of crosslinked elastomers. It is still possible to employ
these results for high Deborah flows of polymer melts when speculatively
utilizing the concept of fluidity loss. Indeed, one can imagine that after
relaxation transition to the high-elastic state, the entanglements in
polymer melts will act as crosslinks in cured elastomers. If the adhesion
strength of polymer to the wall is weaker than that of cohesion, the wall
slip of molten polymer will occur. This seemingly will happen similar to
crosslinked elastomers, i.e. without transport of macromolecules from the
polymer bulk to the sliding interface.
It has been demonstrated for elastomers sliding over a smooth
surface that the dependence of frictional force on velocity shows a
broad maximum. This dependence has been widely observed experimentally529-532 and qualitatively explained by Schallamach533.534 theoretically. Recently, Barquins 535 and Barquins and Roberts 536 reconsidered
anew Grosch's old data 530 and found some quantitative discrepancies,
although the qualitative effects remained the same. This comparison is
shown in Fig. 11.13. 536 It was also found in these works that the
dependence of friction force on the sliding velocity F(v s ), after going
through a maximum, can have a second increasing branch.

-7

-5

-3

-,

,;

Reduced rate. 19 a,V

Fig. 11.13. Coefficient of sliding friction for nitrile rubber pad sliding on wavy
glass (WLF Transform). Results by Grosh 530 for the same rubber are also shown.
The values of sliding velocity, V are (_) 10- 3, (0) 10- 2 , ( ....) 10- 1 and (0)
1.0 mm s -1, respectively.536

384

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

A detailed theory of adhesive friction of elastomers 17o was recently


developed using some of the previous ideas of Schallamach,533,534 which
quantitatively described the effects of velocity on the sliding friction of
rubbers in molecular terms. The theory models the events on friction
contact as stochastic motions of independent macromolecules attached to
the bulk polymer 'body', which are stretching, breaking off and attaching
again to the wall under the action of both thermal excitation and
stretching of individual polymeric links. The theory predicts the first,
increasing branch on the sliding friction characteristics F(v s ) as a result of
the prevailing effect of stretching of the macromolecules on the contact,
and the second, decreasing branch due to their cooperative detachment
from the wall. The theory could not, however, predict the second
increasing branch on the F(v s ). This was accomplished recently in Ref.
171 where the following dependence was obtained:
f(u)=u 1-(1 +m+ l/u) exp( -m-l/u)
1 +ym-exp( -m-1/u)

(11.1)

(Jw being the shear stress at the contact, u = vs/v~. The values of parameters (J~, m, y and v~ were roughly estimated in Refs 170 and 171
through molecular characteristics of the elastomer: at low normal pressures p( P ~ Pm), (J~ ~ GP/Pm, and at high pressures (p ~ Pm), (J~ ~ G. Here G
is the Hookean elastic modulus of elastomer, Pm is a characteristic
pressure depending on the asperities of friction interface. The characteristic sliding velocity v~ is proportional to the Flory radius and reciprocal to
the relaxation time of the part of the macromolecule between two
crosslinks. Parameter y characterizing the effect of Brownian fluctuations
of a thermal segment attached to the wall, is more than unity. Parameter
m, usually very small, is the mean ratio of hovering time of a polymer
bond after its detachment from the wall to the time of bounded state. The
increase in modulus G (e.g. due to growing crosslinking density) results in
a dramatic increase in the value v~ and a substantial increase of
parameter m. At the same time, parameter y decreases slightly.
Figure 11.14171 demonstrates the effects of parameters m and y on
characteristics feu) defined in eqn (11.1). It is seen that the maximum
value of feu) is almost independent of both m and y values. The minimum
value of feu) depends only on parameter y, decreasing when y is growing.
The position of the minimum, Umin, depends on the value of parameter m,
approximately as l/m. Unfortunately, there are no data available with
which the predictions can be carefully compared. The main problem in

385

Melt Flow Instabilities

0.900,...--------,.---------,
I

lb

..

I
I

20 I

0.600

: 2b
I

,,

0.300

0.000 l-~___''--'~___'_ _
-1.500
0.000
1.500
3.000

,,'

,'- "

_ _ L _ ___'_~_:___-'--'

4.500

6.000

Log u

Fig. 11.14. Examples of S-shape dependences of the normalized tangential


stress Qt versus log u (according to eqn (11.1. la: m= 10- 2 ,1'=2'0; Ib: m= 10- 2 ,
1'=5'0; 2a:m=10- 5 , 1'=2'0; 2b:m=1O-S, 1'=5'0. 171

obtaining the data is the onset of either stick-slip on the sliding friction of
the rubber pad, or the occurrence of Schallamach's waves of detachment
when working with spherical sliders. These waves, shown in Fig. 11.15,535
were first described by Schallamach 533 in experiments with elastomers
when the tangential stress on the sliding contact is distributed nonuniformly, as in the case with spherical sliders. This is a typical example
when one has to take into account the bulk mechanical properties of the
elastic body, along with the characteristics of sliding friction. In order to
study these phenomena theoretically or numerically, one has to resort to
solving complete equations of continuum mechanics with some (usually
non-linear) boundary conditions on sliding contact. No such theory for
Schallamach waves has been proposed in literature.
An additional, static friction component of adhesive nature was observed and found equal to 0.063-0.083 MPa in experiments 537 with
peroxided natural rubbers within the restricted time of experiments (24 h).
This critical friction behaviour seems contradictory to the kinetic theories. 170 ,171,533,534 To resolve this contradiction one can notice that in
these theories, the breakage of polymeric bonds was considered both due
to the stochastic thermal excitations, and unconditional adhesive reasons;
the latter occurring when the free energy or elastic force in polymeric
bonds exceeds a critical 'adhesive' value. It means that under any small
given tangential stress, the polymeric bonds are eventually broken. This,
in turn, results in the absence of the adhesive threshold observed in Ref.
537 and extremely slow sliding (perhaps negligible in the experiments)
when the elastic energy accumulated in polymeric bonds is less than the

Fig. 11.15.

Propagation of Schallamach waves in the area of contact of a rigid sphere sliding on a smooth surface of
rubber.535

387

Melt Flow Instabilities

adhesive energy. The situation here is very similar to that well known
for the long-time strength of bulk polymers and other materials. In
this case, more thorough experimental data are also needed. The
static friction component also occurs when there is some surface roughness. In the case of friction of elastomers, however, this component is
small in comparison with that of adhesive friction represented in eqn
(11.1).
A dynamic problem for the simple shear of an elastic plate sliding over
a smooth surface was studied analytically and numerically in Ref. 538
using eqn (11.1) as a characteristic of sliding friction. The exact numerical
results of the self-oscillations, occurring in the unstable region (Ul' U2)
(Fig. 11.16) almost coincided with approximate solution of 'relaxation
type', shown by arrows on the diagram. Additionally, these long periodic
and large-amplitude oscillations were always accompanied by short-scale
and small-amplitude oscillations.
4.0

3.8

3.6

....

f max

fmin
3.4

3.2

-1

2
log u

Fig. 11.16. The phase diagram of inertialess oscillations for sliding friction of
cross-linked polymers 538 and its possible transformations (i)-(iii) for wall slip of
polymer melts: (i) steady slip in the region acb and self-oscillations without
stick-slip; (ii) steady slip region arb and self-oscillations with stick-slip, and (iii)
self-oscillations of stick-slip type without a steady slip region.

388

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

All these features of self-oscillations in sliding friction of crosslinked


elastomers bear a striking resemblance to those found experimentally for
the pulsatile flows of polymer melts. However, to apply these results to
the wall slip of polymer melts, one needs to modify the characteristics
(11.1). The main modification which also might be valid for more robust
modelling of the sliding friction of crosslinked elastomers, is introducing a
static friction component 0";. The modification can be done by 'cutting'
the kinetic sliding friction characteristics, as shown by three vertical
dashed lines in Fig. 11.16. In this case, a new parameter, 0";, the critical
shear stress corresponding to the onset of sliding friction has to be
introduced. It should be noted that the value of 0"; can be taken, as in the
case of vulcanized rubbers, to be ~O.07 MPa, which is in good coincidence as compared with the onset of melt flow instability for the molten
polymers. Depending on the value of wall stress O"w(U2), there can be two
cases: (i) if 0"; < O"w(U2), the self-oscillations in the unstable regime will
happen without the stick condition, i.e. the wall slip will not cease during
the cycle; (ii) if 0"; > O"w(U2), the self-oscillations in the unstable region will
be of the stick-slip type. In both cases, (i) and (ii), there is the stable slip
region, when the sliding velocities are small enough (u < ud. This can
result in the stable steady flow of polymer melt with wall slip, as was
observed by Hatzikiriakos and Dealy.510,511 There is a third possibility,
however: (iii) the sliding friction starts with the decreasing branch of
sliding friction characteristics. This case is related to the occurrence of an
unstable regime of melt flow, accompanied by stick-slip oscillations, as
soon as the adhesion threshold 0"; is overcome. More experimental efforts
are needed to make a final conclusion about the three possibilities.
Additionally, it seems that the possible origin of the sharkskin effects in
the exit capillary zone, is the appearance of the Schallamach waves of
detachment. As mentioned, the waves originate in the regions where the
shear stress is distributed non-uniformly along the sliding contact.

11.5 SELF-OSCILLATIONS OF POLYMER MELTS IN A


RATE-CONTROLLED CAPILLARY RHEOMETER

Lupton and Regester474 derived from their experiments a qualitative


picture of the oscillatory flow of polymers in a rate-controlled capillary

Melt Flow Instabilities

389

viscometer. This instrument consists of two axial symmetrical, successivelyconnected cylinders, a large reservoir of radius R and length L, and a
small capillary of radius r and length l, where 1~ Land r ~ R. The
'integral flow curve' experimentally observed in Ref. 474 is qualitatively
represented in Fig. 11.17. Here
prj(2l) is the wall shear stress,
3
Y= 4Qj(nr ) is the averaged capillary shear rate, p is the pressure in the
reservoir and Q is the flow rate. Usually the length of the decreasing
branch (Yl, Y2) is about one to two decades of Y variation, whereas the interval ('1> '2) is no more than 10% of, variations. Therefore the multiple
region on the integral flow curve often remains unnoticed. The form of
the curve in Fig. 11.17 was shown to be due to the wall slip of polymer,
which at high values of Y contributes overwhelmingly (up to 70-80%) to
the flow rate. 474 In contrast to the opinion of Myerholtz,482 the integral
flow curve in Fig. 11.17 cannot be caused by the decreasing part of actual
flow curve which produces a hysteresis-type integral flow curve with two
increasing branches .

,=

'"1

"

"2

io~

t*

Tz

~ t

Fig. 11.17. A characteristic cycle of selfoscillations in the capillary flow of polymer melts. 539

Tordella 457 pointed out that the integral flow curve in Fig. 11.17 is
almost independent of the hydrostatic pressure. This does not contradict
the sliding friction nature of the curve, because the dependence of friction
force on pressure is caused by the difference between actual and nominal
contact areas, which, in the case of polymer melts, almost coincide.
The cycle of self-oscillations revealed in Ref. 474 consists of the motion
along the contour abcd pointed out by arrows on the phase diagram '-Y
in Fig. 11.17. The oscillations occur only if the mean-integral shear rate
Yu = 4U R2 ja 3 defined by the constant plunger velocity U lies in the
interval (Yl, Y2). For the incompressible material, this can be only if large
cavities are formed in the capillary during the extrusion. A more
reasonable assumption was made and confirmed in Ref. 474 that the
change in Y during the self-oscillations is due to the compressibility of the
polymer in the reservoir. Though at the common pressures of "" 10 MPa

390

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

the volume deformation of polymer in the reservoir is 1-2%, this change


in volume is comparable with the volume of the capillary, which is several
orders less than the reservoir's volume.
In this case, the cycle of oscillations is divided in two stages. On the
first stage, corresponding to the motion 'ab' in Fig. 11.17, the polymer is
compressing in the reservoir and slowly flows into the capillary with or
without small wall slip. This stage is related to the thick portions of the
extrudate exiting the capillary. Then at point 'b' on the diagram the first
slow stage suddenly changes for the second one, rapid flow with the
high wall slip, corresponding to the motion 'cd' on the phase diagram
(Fig. 11.17). At this stage, the shear stress in the capillary is relaxed and
the polymer is decompressing in reservoir. This stage is related to the
thin portion of the extrudate. This continues until the point 'd' on the
diagram is reached, and then the second stage suddenly changes for the
first one. It should be noted that contrary to common opinion, the first
stage of the oscillatory process is not necessarily related to the occurrence of real stick conditions at the wall, and can be accompanied by a
slow wall slip.
The first explanation of the oscillatory flow was given in 1966 by
Pearson (see, e.g. Ref. 2, p. 195). A more complete mathematical treatment
was independently developed in Ref. 539, the calculations from which are
briefly reproduced below.
The approximate mass balance in the system is:

p(t)V(t)~po Vo -Po J: Q(td dt

(11.2)

1,

where
V(t) = Vo -nR 2 Ut+Jp,

p(t) = Po(1 + f3p)

(11.3)

Here V(t) is the actual volume under the plunger in the reservoir, p(t) is
the actual average density of polymer in the reservoir, Po is the polymer's
density at atmospheric pressure, Q(t) is the flow rate in the capillary, 13 is
the polymer's compressibility, J is the compliancy of the reservoir's walls.
Equations (11.2) and (11.3) result in the following:
dr
dt

na 3 ro(Yu

4Vo

-y)+yu r

(11.4)

1 +A- Ut/L

where
r 0 = r /(2113),

A=J/(f3VO )

(11.5)

391

Melt Flow Instabilities

and the very small term 1/3r2 in eqn (11.4) was neglected. If
]iur~ro(Yu -]i), we can also neglect the term ]iur in the numerator of eqn
(11.4). This would give the equation obtained in Ref. 474 on the basis of
which the values of dr/dt were calculated at the beginning of cycles and
coincide with accuracy of 10% with those observed.
Equation (11.4) should be complemented by a rheological equation,
relating the values r and ]i. Within the framework of the averaging assumptions made, such an equation was written in Ref. 539 as
follows:
8(]i) dr/dt + r = cp(y)

(11.6)

The functions 8(y) and cp(y) were simplistically represented on the parts ab
and cd of the integral flow curve in Fig. 11.7 in the piecewisely-linear
form:

(Yo <Y<Yl)
(Ji2 <Y<Y3)

(11. 7)

Finally, the set (11.4) and (11.6) results in the single equation relative the
shear stress r as follows:

11k =

{ '11

112

. {yO
'Yk = .
Y2

(11.8)

(yo <y<yd

(Y2 <Y<Y3)

Equation (11.8) is quickly reduced to a set of recurrent algebraic equations which allow for an exact but somewhat awkward solution. The
solution showed that the period of oscillations decreases linearly with
time, which exactly corresponds to experimental data obtained in Refs
474 and 482.
It should be noted that in common capillary instruments, the
plunger moves extremely slowly. Therefore the period of oscillations
changes very weakly and during an appreciably long time it is almost
constant. Then the expression for the first period of oscillation, depending on the effective shear rate Yu, can be approximately represented in

392

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

the form:

(11.9)

Here TI and Tr are the semi-periods of the first cycle of oscillations


which correspond to the motion of system on the parts ab and cd of
phase diagram, respectively. More precise formulae should be used
instead of eqn (11.9), in the tiny vicinities of values Yo, Yb Y2 and Y3' The
plot of oscillation frequency, OJ = 2n/T1, versus the effective shear rate Yu
is presented in Fig. 11.18. In accordance with eqn (11.9), the ratio TUTr
monotonically decreases with growing Yu, so Tl ~ Tr when Yu is close to
Yi' and Tl ~ Tr when Yu is close to Yz. All the features correspond well to
the data of Refs 474 and 482.
Figure 11.19 demonstrates the change of form for oscillating shear
stress ret) and shear rate yet), depending on value Yu, in the interval of
self-oscillations Y1 < Yu < Y2 If Yu < Y1 or Yu > Y2, the self-oscillations are
w

;
I
I

"

\
I

Fig. 11.18. A qualitative plot of frequency


of oscillations versus effective shear rate
defined by piston's speed as predicted by
eqn (11.9).539

absent. If Yu <Yi or Yu >Y3 then in start-up flow, the function ret)


increasing monotonically tends to a steady value. If however, Yz < Yu < Y3,
then in start-up flow, the function ret) goes through a maximum and then
tends to a steady value. All these features completely correspond to the
experimental data of Refs 474 and 482. Unfortunately, a lack of data
prevented the authors of Ref. 539 from making a quantitative comparison
with data in Refs 474 and 482.
Finally, it is worth mentioning that the same theoretical treatment can
be performed using modern modelling of both viscoelastic and wall slip
phenomena.

393

Melt Flow Instabilities

Or-----~;----r~~----~----

(a)

.r

~~~~__
--.~~-,.---~r-+-,.--~~~
~r-------~----+--r~----~-;---

~r---~-+~--~-r~----;-~-
ro~-=rnr-t~--+-~~~~
O~~--~~--~-r~~~4----

(b)

Fig. 11.19. The qualitative forms of self-oscillations for (a) stress and (b) flow
rate depending on the ratio T UT 1 of semi-periods of oscillations for the first
cycle. 539

11.6 OTHER THEORETICAL MODELS


Attempts were also made to describe the small-scale distortions in
pulsatile flows of polymer melts with the wall slip in capillary and
rotational rheometers without taking into account the effects of polymer
compressibility. As an example of this approach, we briefly consider the
results Ref. 169, where both the viscoelastic properties of polymer in the
bulk and wall sliding friction characteristics were treated as linear to
reveal more easily the basic mechanisms of flow with slip. In this work,
the polymer's viscoelastic properties were represented by a linear threeparametric differential model with relaxation time 0, retardation time Os
(O<s< 1) and Newtonian viscosity 1/. The decreasing linear characteristics of sliding friction, =,o[1-kv(t)J, was also utilized, if the wall shear
stress exceeded the value '0' Also, it was simply proposed that the
systems return to the stick conditions at the wall when the level of shear
stress reached the lower value '* (, * < '0)' Then easy one-dimensional
calculations, with a restructuring velocity profile from bulk flow that also

'w

394

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

involved the wall slip, showed reasonable behaviour of this model. There
is the question, however, of what the real mechanism for the returning
polymer flow from slip to the stick conditions is. This question was
resolved in Ref. 540 where the self-oscillation in Couette flow with wall
slip with decreasing characteristics of sliding friction, were considered for
a non-linear Maxwell liquid. Using the Galerkin method, the problem
was reduced to a set of ordinary differential equations where the inertia
term was essential. Additionally, a two-dimensional instability in plane
flow with fading characteristics of wall slip was discovered in inertialess
flow for Oldroyd B liquid. 54! It is also worth noticing that a twodimensional plane short-wave incremental instability has recently been
found 542 for a problem of flow of an elastic liquid with the sliding friction
at the interface. It was shown that the inclusion of a 'memory term' in the
friction characteristics can destabilize the problem. Thus, in this case, the
instability can also appear on the increasing branch of the sliding friction
characteristics.
There is also the question of how the oscillations in the land region of
capillary flow can be transferred in the extrudate distortions. To some
extent, this question was clarified on the linear model of wall slip!69
where the basic set of equations for the extrudate distortions was
reduced to a set of linear partial differential equations for the recoverable strain and deviations of cross-sectional area and longitudinal velocity in the extrudate from their mean values. The solution of the problem
was obtained in the closed form. Then, using the continuity of elastic
energy flux (10.69) as the matching condition between capillary and
extrudate flows, it was possible to relate the extrudate distortions with
the functions which characterize the oscillatory flow inside the capillary.
It was also shown that the extrudate distortions do occur not immediately but increase downstream from zero to a steady oscillating value
achieved at the distance from die outlet approximately equal to
s8u/(1 + s). Here u is the averaged flow velocity, s is the ratio of
retardation to relaxation, e, times. It is also possible to employ this
procedure for the nonlinear case.
As has been mentioned, attempts were made to describe the spurt flow
phenomena by unstable viscoelastic constitutive equations with stick
conditions at the wall. For example, a good quantitative theoretical
description of spurt flow phenomena and oscillations of Vinogradov's
capillary flow data 292 was achieved in Refs 154 and 155 utilizing the
Johnson-Segalman model (2.10), stabilized by a very small additional
Newtonian term. By fitting Vinogradov's data, the authors of Refs 154

Melt Flow Instabilities

395

and 155 found the parameters of the rheological model. Then in accordance with the common philosophy, they should also have been able to
describe other viscoelastic flows for the same polymers using the constitutive equation with the same rheological parameters. But the authors of
Refs 154 and 155 never attempted to do that. In Ref. 543 the predictions
of the model 154 ,155 were compared with extensional flow data obtained
by Vinogradov and co-workers 507 for the same polyisoprene that was so
successfully used in Refs 154 and 155 for fitting the Vinogradov's
capillary spurt flow data. The results of the comparison 543 clearly
demonstrated that the model 154 ,155 fails to describe in a consistent
fashion the extensional data for polyisoprene in both steady and unsteady
elongational flows.
Further progress in understanding the melt flow instabilities should be
achieved, in the authors' opinion, with more precise experimental investigations, as well as with more profound theoretical and numerical studies
of various types of unstable flows.
(a) The experimental studies of melt flow instabilities should be
focused, in the authors' opinion, on such precise characterizations
of particular types of unstable flow as laser Doppler measurements
of velocity profiles and their fluctuations, combined with the
measurements of stresses by flow birefringence methods.
(b) Several problems of theoretical and numerical modelling are also
important to achieve a deeper understanding of melt flow instabilities:
(i) more precise modelling and experimental verifications of wall
slip phenomena;
(ii) theoretical and numerical modelling of the spurt flows and
self-oscillations related to these;
(iii) numerical modelling of high Deborah entrance flows, especially those related to the occurrence of asymmetrical entrance
vortices;
(iv) numerical modelling of high Deborah exit flows.
Evidently, problems (iii) and (iv) present a challenge to our modern
computational ability.

CHAPTER 12

Additional Problems ln the Rheology of


Polymeric Fluids

This final chapter briefly discusses a number of phenomena observed in


shear and extensional flows of polymeric fluids which, except for those
described in Section 12.3, are at present either not completely investigated
experimentally or poorly understood theoretically.
12.1 STRENGTH OF POLYMER MELTS UNDER EXTENSION
The polymer melts, mainly with narrow MWD, can be ruptured under
intense elongation, even in the range of temperatures high above Tg
These remarkable phenomena were reported by Vinogradov and coworkers with experiments on monodisperse polybutadienes (PB) and
polyisoprenes (PI) in Refs 281, 503 and 528 and clearly confirmed the
concept of fluidity loss as a transition from a liquid to a solid state(s)
under extensive regimes of deformations.
Consider for example, the ruptures arising in extensional experiments
with constant strain rate K, for polydisperse 1.2 PB with molecular mass
M ~ 10 5 at 10 dc. 528 The plots of total stress (Y versus total strain
(defined by eqn (4.38)) are shown in Fig. 12.1. Open circles in the figure
correspond to the rupture of specimens. It is seen that at low value of K
the steady extensional flow is attainable (curve 1). With further growth of
K a rupture with torn edges occurs (curve 2). When the extensional rate

Additional Problems in the Rheology of Polymeric Fluids

397

Fig. 12.1. Stretching stress (J versus full strain


In 8 for polybutadiene at a constant strain rate
K at 100e. Curves 1, 2, 3, and 4 correspond to
K=9.5 x 10- 4 , 6.5 x 10-3, 3.5 X 10- 2 , and
1.2 x 10- 1 s-l, respectively.528

K is larger, the plot (J versus In e has a maximum (curve 3) corresponding


to a neck propagated along the specimen. With still more intense
extension (curve 4), a sample was ruptured by brittle mechanism without
necking; the surface of rupture being as smooth as a mirror. In these
experiments, it was also observed that in both 'necking' and 'brittle' cases
of ruptures, the strain in the samples was practically completely elastic
(recoverable) as for the crosslinked rubbers. Similar results were also
obtained in Ref. 528 at a fixed strain rate with varying temperature. These
results were unified in Ref. 528 in a master curve. Such a master curve in
the dependence of effective elongation viscosity (J/(K()) on the dimensionless time t/() is shown in Fig. 12.2.582 It is remarkable that the viscoelastic
behaviour with relaxation transitions to high elastic and even glassy state,
observed in these experiments, holds until the rupture. This means that
there is no evidence of thermo-activated Euring's fracture mechanism.
Similar observations, but sometimes more blurred, have been reported
in the literature with rupture of polydisperse polymers. Thus for the
extension of polyethylene,276 little pronounced neck formation with
subsequent rupture of sample was observed.
It has been shown 528 that the dependence of stress at the instant of
rupture or necking, uP' on recovery strain IX is independent of the uniaxial
extension regime. Also, the plot (Jp versus In IX can be presented as a linear

398

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

-I

Log f (5)

Fig. 12.2. The temperature-invariant relation between the reduced extension


time and the true stress normalized with respect to viscosity and rate of extension
deformation. The temperature of polyisoprenes of molecular mass 5.75 x 10 5
varied from Q to 75C: 0, QOC; .6., 25C; D, 5QoC; <), 75c. 281
dependence which is very characteristic for crosslinked rubbers. This plot is
demonstrated in Fig. 12.3 for mono disperse 1.2 PB. The minimum elastic
strain In & for which rupture was recorded was equal to 0.5. Detailed studies
showed that for monodisperse polymers, the dependence O'p (In &) is
independent of temperature and molecular weight of linear polymer.
The similar dependence O'p{ln &) will apparently be different for polydisperse polymers. As an example, consider the extension of PIB at
K= 3.84 x 10- 3 shown in Fig. 6.4, where contrary to mono disperse
polymers, steady extensional flow occurred at &~ 1.9 (In oc~0.642). It

1.0

o.s

2.0

-(n~

Fig. 12.3. Plots of rupture stress (5 versus elastic strain In eX. Different points correspond to the regimes of extension at
constant strain rates at stress at various
temperatures. 582

Additional Problems in the Rheology of Polymeric Fluids

399

should be also noted that Figs 6.9 and 6.10 demonstrate the existence of
envelope for the dependence <T(&), to the right of which there is a
forbidden region, non-attainable for any regime of extension. Along with
the rupture dependence <Tp(&), the envelope defines the maximum values
for elastic strains attainable at various stresses.
There are two more curious observations found by Vinogradov and
co-workers for the monodisperse polybutadienes and polyisoprenes,
which are in striking resemblance with the strength behaviour of cured.
rubbers. The first is the failure envelope shown schematically in Fig.
12.4 281 for these polymers, which is well known for the cured rubbers as
the 'Smith's envelope'.544 The envelope splits the total <T(e) diagram in
two (forbidden and attainable for deformation) zones; the marginal line
being the line of rupture. The lower part of the marginal line relates to
relaxation transition to rubbery state when extension rates are relatively
small, the upper part of the line, to relaxation transition to the glassy
state under action of high strain rates. One can find in Fig. 9 of Ref. 503
an example of constructing the Smith's envelope from experimental data.
The second observation relates to the long-term durability of uncured
polymers. This is demonstrated in Fig. 12.5 281 as WLF plots for monodisperse polybutadienes and polyisoprenes. For both polymers, these
plots present some power-like dependences with almost the same value of

10.---------------

0.1

1
Fig. 12.4.

E*

A diagram of the failure envelope for incured high molecular linear


polymers of narrow MWD.2B1

400

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

J
.......
~

... 1

I-'
\::)

Cl

.Q

log (JI~ (MN m- 2 )

Fig. 12.5. The long-term durability plots for (A) polybutadienes and (B) polyisoprenes, invariant relative to temperature and molecular mass, for different
molecular weights. 2s1

power exponents. It is also interesting that the ultimate elastic strains


before rupture can be extremely large (~10-15). These findings also
bear a striking resemblance to the findings for uncured rubbers.
Vinogradov 281 argues that the durability plots indicate the mechanism of
rupture associated with accumulation of defects.
12.2 ON DETACHMENT OF POLYMER FROM WALLS
12.2.1 Detachment of Polymer from Capillary Wall under
Stretching of Extrudate: Draw Resonance

Stretching a polymer extrudate exiting the capillary was considered in


Section 10.7. The experimental scheme was shown in Fig. 4.17. Contrary
to the free polymer extrusion from a capillary die, where the radial stress
presses the polymer against the wall, the extrudate extension can change
the sign of radial stress at the wall in the exit die region, causing the
detachment of polymer from the wall. This effect is well known in

Additional Problems in the Rheology of Polymeric Fluids

401

industrial fibre spinning and has been reported in the literature several
times for various polymer systems: polyesters,545 aqueous solution of
polyethylene oxide,436 polyisobutylene melt,433 and recently for aqueous
glycerin solutions of Separan AP30 of different concentrations. 546 The
effect was most intensively studied in Ref. 433 where it was also shown
that the onset of the detachment depends not only on the drawing ratio
and rheological properties of polymer fluid, but also on the wall's
material. The latter is attributed to failure of adherence of polymeric
liquid to the wall, which has not yet been studied quantitatively, either
experimentally or theoretically. Sometimes the effect is accompanied with
pulsations of the stretched extrudate's diameter, as has been observed in
Ref. 436.
A periodic change in extrudate diameter with time under its stretching,
without detachment of polymer from the die walls, has been observed
many times in the literature. These oscillations were called 'draw resonance', because they begin, even under isothermal conditions, at a certain
drawing ratio. 54 7 A good review of the effect can be found in Ref. 462.
The draw resonance was observed both for the Newtonian 196 and
viscoelastic fluids.548-551 Both linear 547 ,549,55o,552,553 and recent nonlinear 554 ,555 theoretical analyses showed that the elasticity of liquid plays
a secondary role on the onset of the oscillations.
12.2.2 Detachment of a Deformed Polymer from the Wall
under the Action of a Low Molecular Fluid384

These phenomena were observed in the course of experiments 384 ,386


discussed in Section 9.7 There, shearing of a polyisobutylene of
MW", 104 between rotating disks was. used in a shaft-sealing device to
separate water from air (see Fig. 12.6). It was noticed that sometimes
water placed above the polymer layer (shaded area) penetrated through
the layer, between the polymer and the steel wall.
In these experiments, which were carried out at room temperature, the
polymer was placed into two gaps between the mobile disks (1 and 2) and
the immobile disk (3). The mobile disks were rigidly connected with a
rotating shaft (4), the immobile one, with the cylinder (5); the two
distances between mobile and immobile disks being equal. The shear rate
y at the external edge of disks was estimated as '" 102 S - 1.
The advance of water film penetrating through the polymer layer did
not depend on the direction and value of stresses arising between the
disks. The penetration of water occurred only when the polymer was
deformed and looked as the polymer was detached from the wall. Also,

402

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

H
H
_l_

Air

..L
][

Fig. 12.6. Sketch of a device for studying


the detachment of sheared polymer from the
wall by a water film. 1, 2, Mobile disks; 3,
immobile disk; 4, rotating shaft; 5, cylinder.

when the water was replaced by air, no penetration of air through the
polymer layer, as well as the polymer detachment (at least, complete)
from the wall was observed. By coating the steel wall by acrylic plastic,
the penetration of water through the polymer layer disappeared, even in
shearing. This effect is apparently due to the surface tension acting at the
interface between the water, the wall and the polymeric liquid.

12.3 CAPILLARY BREAK-UP OF POLYMERIC JETS


Unlike the other sections presented in this chapter, the phenomena
consider here are well understood both experimentally and theoretically.
The capillary break-up of Newtonian jets freely flowing out of the
nozzle is a well elaborated part of capillary fluid dynamics, originating
from the work of such scientists as Raleigh and Bor. A comprehensive
review of results in this field can be found in Ref. 556, pages 123-8. Figure
12.7(a), taken from Ref. 557, demonstrates a typical capillary decay of
horizontal jets of Newtonian liquids with the formation of free droplets.
Surprisingly enough, linear analysis of infinitesimal disturbances predicts
very well such basic characteristics of the break-up as the length I until
decay, and the size of droplets.
New phenomena in the capillary break-up of jets of viscoelastic fluids,
mostly 0.25% aqueous solutions of polyacrylamide, were discovered in
Ref. 557. In accordance with elementary analysis, the disturbances in
Na-CMC solution initially grew more rapidly than in a Newtonian fluid
of the same viscosity. But after reaching a certain amplitude, the grow of

403

Additional Problems in the Rheology of Polymeric Fluids

I
II1II1111.II1II

.11111.

(a)

tI

Fig. 12.7. Capillary decay of jets of a Newtonian and polymeric liquids: (a) 70%
glycerin in water; (b) 025% aqueous solution of Na-CMC; (c) 025% aqueous
solutioh of polyacrylamide. 557

the disturbances was sharply slowed down and, before the rupture, a line
of small droplets connected by very thin threads was formed (Fig.
12.7(b)). For the polyacrylamide solution, more elastic than the solutions
of Na-CMC, no wave formation was observed and the first visible
disturbance appeared in the form of a large separated drop connected by
thin threads of random length (Fig. 12.7(c)). Some intermediate structures
were also observed in Ref. 557 for other elastic liquids.
All viscoelastic fluids studied in Ref. 557 displayed the rupture distance
1 to be much larger than that corresponding to Newtonian liquids.
Similar but less representative results were obtained with low-concentration polymer solutions. 558 Only PIB solution in tetraline 559 was found to
follow the predictions of linear analysis of viscoelastic disturbances with
value lless than in the corresponding Newtonian liquid.
The mechanism of capillary break-up in polymeric jets was qualitatively explained in Ref. 560 where the authors pointed out that the dropletthread structure is caused by stretching of the threads when the droplets
are forming under the action of surface tension. One can assume that a
sharp loss of fluidity of an elastic liquid under extension, accounts for the
surprising stability of threads between the droplets. Also, in the short
nozzles, the jet can be appreciably stretched when exiting the capillary.
This can also slow down the growth of capillary disturbances and lead to
an increase in the length 1.
These ideas were employed in theoretical calculations 561 based on
quasi-one-dimensional jet dynamics and a molecular rheological
mode1 562 which predicts a sharp increase in extensional viscosity. The
authors of Ref. 561 succeeded in describing analytically and numerically

404

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

the process of jet evolution with the quasi-steady droplet-thread structures, assumed to be formed according to experimental observations. But
the mechanism of formation of the structure remained unclarified. These
results were later reproduced numerically in Ref. 563, where a Maxwellian rheological model was used.
Also, the capillary break-up of vertical jets of aqueous polyacrylamide
solutions was studied in Refs 564 and 565, where the same effects were
observed. An additional extension under the action of weight can cause an
even higher stability of jet. But this effect was not verified experimentally.

12.4 FLOW-INDUCED CRYSTALLIZATION


In 1949 Kargin and Sogolova 184 ,185 started to study simple extension of
high viscosity ('7 '" 109 Pa s -1) polyisobutylene at a constant stretching
force. They found that at certain values of force the flow started to slow
down. Later studies 283 ,566 where X-ray and electron microscopy analyses
were used, showed that the PIB crystallized in flow. Figure 12.8 demonstrates the crystallization of the PIB in a long time experiment,566 lasting
several days. The upper, saturating branch of the curve is related to
crystallization. In contrast to the stretching of LDPE (Fig. 7.6), no repeated
development of flow was observed. The crystallization disappeared when

2B
0

0
0
W

20

12

4
0

16

48

80
t

112

144

176

(h)

Fig. 12.8. Total strain versus time t (h) in stretching high molecular polyisobutylene under a constant force. 566

Additional Problems in the Rheology oj Polymeric Fluids

405

loading was removed. Also, some crystallization effects were observed 567 in
elongational flows of polypropylene in extensional rheometer of Meixner
type. At suitable temperatures, certain critical elongation rates were
detected in these experiments, above which the crystallized spots along the
extended fibre occurred. Flow-induced crystallization was also reported in
Ref. 568 for an isothermal flow of polypropylene in the entrance capillary
region, where the extensional flow is dominant. Even more remarkable
effects have been found 569 in an isothermal polymer flow in the entrance
region of an annular die, where the flow was nearly extensional. At some
temperature, the crystallization in flow occurred in the form of a continuous
fibre which could be detected even in the extrudate.
In contrast to the simple elongation, the effects of flow-induced
crystallization in simple shearing are not so pronounced and usually
occur in high shear rate flows, but they are easily detectable. Flowinduced crystallization in capillary flows of polymer melts was observed in
Ref. 570. A lot of experimental evidence reviewed in Ref. 567 exposed the
crystallization in shear flow very distinctly. These effects were observed
under steady conditions by birefringence measurements coupled with
rheometry. In particular, it was found that after an induction period, the
shear stress sharply increased, especially under cooling conditions. It is
remarkable that the rate of crystallization in a flow can be several decades
higher than that in the quiescent crystallization of polymers. This is
explained in Ref. 567 by the highly increased speed of nucleation in flow.
A kinetic model for crystallization in steady shearing flows of polymers
proposed by Janeschitz-Kriegl and co-workers has been discussed in
detail in many publications (for a review, see Ref. 567) and seems to be
robust and descriptive. But in order to be more applicable to various
processing operations, the model needs to be extended to unsteady and
general types of polymeric flows. The next necessary stage in modelling
these phenomena is a description of the rheological behaviour of polymer
melts during crystallization under non-isothermal quiescent and flow
conditions. No studies in this direction have been carried out.

12.5 POLYMER DEGRADATION


Polymer degradation is defined as a combination of chemical and
physical changes occurring during the processing, storage and use of
polymer materials and resulting in the loss of some useful properties of
these materials.

406

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

The degradation of polymers is more complex than the depolymerization on its low molecular weight compounds of similar structure. This is
due to the fact that the polymer structure is not simply a repetition of
polymer units. Such polymer features as chain branches, unsaturated
links, impurities and a variety of other factors impose drastic effects on
the degradation phenomena.
Because of the importance of degradation for all polymer technologies
and applications, many papers and books have considered the problem.
In the following brief discussion, we only refer to Ref. 288 and to two
recent monographs 571 ,572 devoted to the chemical and physical aspects
of polymer degradation.
There are many types of polymer degradation: thermal, photo, oxidation, mechanical degradation and oxidation, degradation caused by
electric fields, etc. From the viewpoint of polymer rheology and processing, the most important of these are the thermal and mechanical
degradation and oxidation.
12.5.1 Thermal Degradation and Oxidation

The complicated nonlinear viscoelastic properties of polymer melts and


solutions are not so much welcomed in polymer processing operations
and rheology, because they usually create limits in increasing productivity. Since the relaxation phenomena, which are notoriously limiting the
throughput, sharply decrease with increasing temperature, it is natural to
establish the regimes in processing operations with temperatures as high
as possible. The thermal degradation and oxidation phenomena, whose
rates also increase sharply with increasing temperature, impose, however,
additional constraints on the upper values of the temperatures attainable
in polymer processing and rheological measurements. These reactions
often result in a substantial decrease of polymer molecular weight by
chain scission.
There are two main classes of thermal degradation, related to depolymerization and substitution reactions. 5 71 In the first case, breakage
of the backbone occurs, while in the second case, the main character of
repeating units but not the length and the character of the main chain, is
changed.
Although all polymers degrade at high temperatures in the absence of
oxygen, degradation is almost always faster in the presence of it. The
reaction of oxidation of hydrocarbons is normally of an auto-accelerating
type, with a slow or even negligible first stage followed by a noticeable
increase in rate.

Additional Problems in the Rheology of Polymeric Fluids

407

Many elementary chemical reactions in the thermal and oxidative degradation of polymers are accompanied by a change in the MW and MWD
of their macromolecules. Among these reactions are the decomposition of
macromolecules and macro radicals, and reactions of recombinations. The
former result in a decrease of MW, while the latter result in an increase. The
presence of oxygen shifts the reactions to a decrease of MW and MWD.

12.5.2 Mechanical Degradation and Oxidation


During many melt processing operations, considerable extension and
shear stresses are applied to polymer melts leading to the formation of
macro radicals. Though infrequently, oxygen can efficiently react with
these radicals, resulting in a permanent break of polymeric chains.
Rubber processing commonly uses high stresses which lead to a much
faster rate of degradation than for polymer melts. This usually is
exacerbated by the presence of ozone in the atmosphere.
The rupture phenomena for monodisperse polybutadienes and polyisoprenes discussed in Section 12.1 can be considered as the manifestation
of mechanodegradation on the macroscopic level.
The mechanical degradation may also be responsible for the anomalous behaviour of LDPE in highly intense elongation and shear flows
discussed in Sections 7.2, 8.6.1 and Chapter 11 of the book. It should be
remembered that contrary to HDPE, the elongational viscosity for
LDPE goes from the maximum. This behaviour has a tremendous effect
on the capillary entrance flows at high Deborah numbers and on the melt
flow instabilities. However, the question is whether this is a true rheological effect or it is related to mechanical degradation. Some experimental
evidence discussed in Chapter 7 demonstrated that the repeated extensional experiments resulted in a small but noticeable decrease of elongational viscosity. Though there are pure theological means to describe this
effect, the problem is that, unlike mechanodegradation, the rheological
descriptions proposed in the literature are essentially reversible, meaning
repeating loading and unloading. There is a crucial need for experiments
to resolve this contradictory situation.
12.6 PROBLEMS IN THE RHEOLOGY OF OTHER
POLYMERIC SYSTEMS
Although throughout this book, we have considered only rheological
properties such 'pure' polymeric systems as long chain flexible polymers

408

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

and their solutions, many studies, especially in the last decade, have been
devoted to other, more complex polymeric systems of industrial importance.
12.6.1 Rheology of Filled Polymers
Filled polymers form a vast class of complex systems used traditionally
in the rubber and tyre industries. The common compounds consist of
natural or synthetic rubbers highly filled with black carbon. Small
(usually submicron) filler particles with additional attractive interparticle
or polymer-particle interactions, create a particulate labile network
which is ruptured in flow and restored in the rest state. In slow flows,
these peculiarities of structure usually result in the occurrence of yield
and thixotropic behaviour. Though these systems are of great industrial
importance, there is still a need for extensive experimental and theoretical studies of their rheology, especially for modelling their rheology and
industrial processes. Two types of continuum mechanics (phenomenological) descriptions were developed to model the rheological properties of the systems. One of them, which uses the traditional approach
of the plasticity theory with yield criteria modified by a viscoelastic
functional, was recently reviewed in Ref. 573. Another approach 214
operates with more detailed kinetic description of the labile particulate
network, additionally to the viscoelastic constitutive equations. This
approach predicts the yield and thixotropic phenomena without mathematically complicated yield criteria, as a transition of a bifurcation
type from static to flow behaviour. Though this approach is very
promising, especially in the applications, its testing was started very
recently. Some more detailed micromechanical theories reviewed in Ref.
214 have not resulted yet in the constitutive equations consistent with
experimental data.
12.6.2 Rheology of Polymer Blends
This quickly developing field of study is very important for industrial
applications. Most studies are industry-oriented and focus on how to
create polymer blends and compounds with material properties better
than the properties of their constitutents. The blends can be thermodynamically compatible or not. In the first case, some physical
approaches can be applied to study thermodynamic and kinetic conditions of compatibility and phase separation. The problem, however, is
that these approaches are usually not applicable to the flow phenomena
that occur in processing operations of polymer blends. Some compli-

Additional Problems in the Rheology of Polymeric Fluids

409

cated rheological properties occur in flows of incompatible polymeric


blends, too. These are related to the multiphase character of flow with
specific surface interactions between components of blends. Many results were recently reviewed in Ref. 574.
12.6.3 Rheology of Liquid Crystalline Polymers
This is also a rapidly developing field of study, although without so
many industrial applications expected in the near future. The physics
of the systems with rod-like macromolecules, as well as their rheology,
has had a great impact on recent studies (see, e.g. Ref. 6, Chapter 10,
and Ref. 575). It was understood that dominant dynamic effects in
the concentrated polymeric liquid crystals are mostly caused by topological constraints. These systems display highly anisotropic behaviour
in flow, and a domain or 'mosaic' structure in the equilibrium at rest.
This rest structure, usually unknown in detail, greatly affects the flow
of liquid polymeric crystals, and creates some uncertainties in their
rheology.
12.6.4 Chemorheology
Despite numerous industrial applications, this field of study is developing slowly, seemingly because of many experimental difficulties in obtaining reliable rheological data. The main problem of chermorheology
is the study of rheological properties of systems during the polymerization or curing reactions. Of the most popular type of reactions (chain
grow, step grow and homo- (or random) polymerizations), the steepest
increase in the rheological properties occurs in the random polymerization near the gelation point. The measurements of rheological properties should include their dependence on reaction conversion, strain rates
and temperature. The latter is a particularly difficult task because of the
high rates of common reactions and the need to separate the temperature dependences of reactivities from these for rheological properties
(viscosities and relaxation times). Such experiments can be performed by
instant 'cooling' the reaction to the region of temperatures where the
reaction rate is negligible but flow still exists. 576 Then rheological
measurements of the cooled reaction mixture with a fixed conversion at
different temperatures (within the above temperature region), can evaluate the dependence of rheological parameters on the conversion, including the temperature dependences of characteristic activation energies of
flow. Some useful empirical correlations for Newtonian viscosity are
reported in the literature (see, e.g. Chapter 5 of Ref. 577).

Appendices

Al KINEMATICS OF CONTINUUM
This appendix presents only that material which is directly used in the
basic text. The reader is assumed to be familiar with elements of tensor
algebra and analysis that can be found elsewhere (see, e.g. Refs 578 and
579). For more details on the common presentation of continuum
kinematics see Refs 94, 148, 371 and 580; and for more strict results of
rational mechanics, see Ref. 93.
A1.1 Eulerian and Lagrangian Descriptions

Let Ox! X 2 X 3 be a fixed coordinate system relative to which a motion of


continuum is studied. Along with a clock it forms a fixed frame of reference.
In the vicinity of any fixed point x we can imagine a small 'volume of observation' surrounding the point, through which there is a 'flow' of material
points of continuum. The method describing the motion of continuum from
the viewpoint of a fixed observer is called the Eulerian description.
There is also a dual method describing the motion of continuum
similar to that accepted in theoretical mechanics for a set of material
points. Consider, for instance, a set of coordinate lines, Xi imposed on an
initially immovable continuum and attached when it moves to the same
material points. Then in the motion, these coordinate lines 'frozen' into
continuum will be transformed due to its deformation to the system

Appendices

411

O~1~2C which is generally very different from the initial one. The
coordinates ~i of material points of the continuum are called Lagrangian
coordinates. Along with a clock, these form a movable (deformed and
rotated generally) frame of reference. In Newtonian mechanics, the time
in movable and fixed frames of references is the same.
The coordinate lines ~i are demonstrated in Fig. AI, where for the sake of
simplicity, a two-dimensional picture is shown. Usually it is also assumed
that initially ~i = Xi. The method describing motions of continuum by
using Lagrangian coordinates is called the Lagrangian description.
All the information about the continuous deformation of a medium is
contained in the continuum law of motion, i.e.

x=x(r,l;)

(O~r~t),

x(O, 1;) = I;

(A.I)

which is assumed to establish at any time r a smooth one-to-one


relationship between the Eulerian (x) and Lagrangian (1;) coordinates of a
continuum.

{x(~,O)

e2

o
Fig. AI.

Motion of an element of continuum.

Though both descriptions are identical, the relations between these are
complicated and generally implicit. Thus the choice between the two
descriptions is usually made on the basis of simplicity. For instance, when
studying the deformation in elastic solids, where the distance between two
arbitrary points is limited, it is more convenient to use the Lagrangian
description. On the contrary, when studying the flows of liquids, and
especially their steady-state flows, it is more convenient to use the
Eulerian description, as is done in this book, which is why the Eulerian
description is developed in the following parts of the appendix.

412

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

A1.2 Basis Vectors and Tensors


In contrast to common tensor analysis, all the peculiarities in the

definition of tensors in continuum mechanics are related to the fact that


the basis vectors are generally mobile. To clarify this, let us introduce
some useful definitions.
The motion of a certain material point M in continuum is described by
the movable radius-vector r(,,~) originating at the fixed point O. At any
time, in the past we can define three basis vectors ei crossing the material
point M and being directed along the tangents to the coordinate lines ~i,
as follows:
(A.2)

Then using eqn (A.2) we can define two main sets of basis vectors, related
to the initial (,=0) and actual (,=t) time instants:
(A.3)
Here the first relation in eqn (A.3) is valid due to the initial condition in
eqn (A.t). Also, at the time instant, = t, the material point M with fixed
coordinates ~ is related to the basis vectors
(A.4)

If in the fixed frame of reference, the system of coordinates OX 1 X 2 X 3 is


Cartesian, the basis vectors ei are independent of space coordinates and
time, and ei == ei'
Having defined the basis vectors in both frames of reference, it is easy
to define the metric properties of these by introducing the covariant
components of metric tensors:
(i, k= 1, 2,3)

(A.5)

Here the overcircles denote the scalar product of two vectors. Then the
contravariant components of the metric tensors can be also determined
from the common relations:

b~={l, i=k
0,

i =fk

(A.6)

and the vectors of dual basis are defined as:


(A.7)

Appendices

413

Hereafter the common rule of summation by two repeating indices is


assumed. From eqns (A.6) and (A. 7), some additional useful relations can
be obtained. For instance,
(A.8)
By using the relations (A.5)-(A.8) we can now describe any components
of a tensor and find a convenient polyadic representation for any tensoric
object. For example, the metric tensor b can be described by the diadic
repesentations as follows:

b = gik eiek = gikeie k= b~eiek


(A. 9)
Inasmuch as the same way we can represent, e.g. a second-rank tensor A
as:
(A.IO)
where the basis vector in eqn (A.IO) are either overcircled or capped
depending on the frame of reference in which the tensor A is defined. Now
all common tensorial operations can be defined in the frames of reference.
At.3 Strain Gradient Tensors
Consider now the local properties of the continuum law of motion (A. 1).
Due to the smoothness of the one-to-one dependence (eqn (A. I)) assumed,
at any actual time t there exist two dual local linear transformations
between the Lagrangian I; and Eulerian x coordinates:

(A.ll)

Here, for the sake of simplicity, the vector differentials of spatial coordinates at constant actual time t, bl; and bx are denoted as follows:

bl; = dr(O, 1;),

bx=dr(t,l;)

(A.12)

The determinants of these transformations do not vanish and are related


as:
det F= l/det 4>

(A. 13)

According to eqns (A.ll), the mutually inverse second-rank tensors F and


4> can be represented as corresponding gradients:
4>=V~x

(A.14)

414

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

which is why both of them (more often tensor F) are known as strain
gradient tensors. In coordinate form, the tensors F and can be expressed
as follows:
(A.1S)
Equations (A.ll) can be also written as the transformations between the
two basis vectors, Cj and ej:
(A. 16)

Here tensors F and are considered as linear operators in [R3. Due to


the one-to-one smooth dependence (eqn (A.1)), the tensor gradients F
and can depend either on ~ and t or on x and t variables. Tensor F is
more convenient to use for Lagrangian descriptions where usually
F = F(t, ~), whereas tensor gradient is more convenient for Eulerian
description and then = (t, x). In the following much more attention is
paid to the Eulerian description, i.e. tensor gradient (t, x) used earlier
in this book.

Al.4 Cayley Polar Decomposition


Cayley polar decomposition in application to tensor gradient states
that it can be represented as follows:
(A.17)
Here R is the orthogonal tensor and i and), are symmetrical ('left' and
'right', respectively) strain tensors. With few very artificial exceptions, the
tensors i and ), can be considered as positive definite. These tensors in
Eulerian description are functions of time t and spatial coordinates x.
Also, the dual polar Cayley decomposition for tensor F is written in the
form:

F=VR*=R*U

(A.18)

where V and U are left and right strain tensors (symmetrical and almost
always positive definite second-rank tensors) and R* is another orthogonal tensor. Because the gradient tensor F is inverse to the tensor ,
there are relations between orthogonal tensors Rand R* as well as
between the Eulerian strain tensors i and i, and Lagrangian strain
tensors V and U:
(A.19)

Appendices

415

At.5 Strain Measures


In accordance with the first relation in eqn (A. 11), the square of
infinitesimal distance between two close points of continuum at initial
time instant t = 0 is given by:

(A.20)
where the symbol
defined as follows

means the operation of transposition. The tensor C


(A.21)

is called the Green (or sometimes the Almansi) strain tensor. The last
relation in eqn (A.21) is due to eqn (A.18). Sometimes, instead of the
Green strain tensor, the Green measure of strain G is used. This is defined
as follows:
G=l(o-C)

(A.22)

Equation (A.20) and (A.21) show that the Green tensor C is the image of
the metric tensor b in the basis ei where t\ = gij. In the basis ('ii, the
components of the Green tensor are the following:
_
8~m 8~k
Cij = 8x i 8x j gkm
If we consider the square of infinitesimal distance between two close
points of continuum at actual time distant 1: = t, then similar to eqn
(A. 20), the following relations hold:

(dl)2 =bx' bx=b~ .ljJt.ljJ. b~

(A.20a)

where the tensor C- 1 defined as follows


(A.21a)
is called Finger strain tensor. It is also easy to see that the Finger strain
tensor is the image of metric tensor b in the basis ei where (C-1)ij=gij.
In contrast to the strain gradients F and ljJ, strain tensors C and C- 1
provide us only with information of pure deformations of a medium; the
local information of rotations of infinitesimal particles being lost. That
is why the Green and Finger tensors are named strain tensors. Nevertheless, if the total field of deformation throughout a deformed body is
given, it is also possible to restore the rotations, provided there are no
cracks and voids in the continuum. The latter condition results in the
fact that the deformed body is still a part of Eucledian space whose

416

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

curvature tensor formed from space derivatives of tensor C (or C- 1 ) is


equal to zero. These are important conditions of compatibility for the
continuum deformation.
Having defined the Green and Finger strain tensors, we can also define
other measures of deformation useful in applications. Let us consider a
set of symmetrical second-rank tensors in [R3 which generally depend on
space variables and time. If the value of a tensor T (i.e. its components in
any basis) is completely determined by the values of metric tensor g and a
tensor H, the tensor T is called isotropic tensor function of tensor H. This
dependence is written as T = f(H). There are two important properties of
isotropic tensor functions:
(i) tensor T has the same principal axes as tensor H,
(ii) the principal values Tk and H k of corresponding tensors are
related as follows:
(k= 1, 2, 3)

(A.23)

Usually it is assumed that the isotropic tensor function is generated by an


analytical scalar function f(z). Ifthe function fez) is strictly monotonous,
then the isotropic tensor function T= f(H) is said to be 'monotonous'.
For instance, the isotropic tensor function T=exp H is defined as usual
power series of exponential function:
T=fJ+H+!!H2+t!H3+ ...

and is monotonous.
Now we define a strain measure as any isotropic monotonous tensor
function of Finger (or Green) strain tensor C- 1 ( or C). This means that
these strain tensors are strain measures. Another example of Green strain
measure G has been introduced by eqn (A. 22). We can introduce one
more strain measure, called the Hencky strain measure h, as follows:
h =! In(C- 1 )=!(C- 1 _ fJ)-t(C- 1-fJ)2 + ...

C- 1 =exp(2h)=fJ+2h+2h2 + ...

(A. 24)

Equations (A.24) mean that h ~ 0 when C-1_-+fJ, and moreover, it is well


known that h~1'. when h~O. Here I'. is common infinitesimal strain tensor.
A1.6 Invariants of Tensors and Hamilton-Cayley
Identity: Density
Though tensors are invariant geometric objects, their components change
with changing coordinate systems. Still there are some scalar functions of

417

Appendices

tensor components that do not change under arbitrary transformations


of coordinate systems. These are called the scalar invariants. There exist
an unlimited amount of invariants related to any tensor. But for any
symmetric second rank tensor c in 1R3 under consideration, there are only
three independent scalar invariants. Usually it is convenient to operate
with three scalar invariants 1 1 , 12 ,I 3 which appear in the well-known
Hamilton-Cayley identity:
(A.25)
Here

13=detc

11 =trc,

(A.26)

In a curvilinear coordinate system, the trace operation is defined in the


common way:

trc= cikgik =Cikg ik = c~


and the determinant means detci. Sometimes, other invariants of a tensor
c can be also used. For instance, the principal values Ck of tensor care
useful. The basic invariants Ik of tensor c can be represented through its
principal values, Ck as follows:
It is worth noticing that due to expressions (A.27), the principal values

Ck

also satisfy the Hamilton-Cayley identity (A.25). Also, because of the


Hamilton-Cayley identity, any isotropic tensor function of tensor c, t/I(c),
has the general representation:
(A.28a)
Here the scalars kj are functions of three independent invariants of tensor
c, say I k If, additionally, the tensor c is positive definite, the dual general
representation is:

t/I(c) = moo +mlc+ m2c-1

(A.28b)

Equations (A.28) can, in principle, be obtained by expanding the regular


function t/I(c) in Taylor series and then using the Hamilton-Cayley
identity for the terms of expansion with power equal or higher than three.
Until now, tensor c was an arbitrary second-rank symmetric tensor.
Consider now the case when e is the Finger tensor, i.e. e = C- 1 . Then e is
positive definite by definition. Let us consider the physical sense of
invariant 13 = dete. Without loss of generality, we can choose at r = 0 the

418

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

initial coordinate system {~} as Cartesian, and the coordinate system {x}
at actual time T = t as orthogonal (generally curvilinear) and coinciding
with that formed by principal axes of tensor c. Consider now infinitesimal
parallelepiped in initial state formed by the componets d~i' Its volume is
dVo=d~1 d~2d~3' The image of the initial elementary parallelepiped in
the deformed state is formed by the components dXi = Ai d~i = d/2d~i'
Thus its volume is
dV(t) =dx1 dX2 dX3 =(C1C2C3)1/2 d~ 1 d~2 d~3 =g/2 d Vo

(A.29)

By their construction, the elementary volumes d V(t) and d Vo have the


same mass, dm. Thus the law of the mass conservation reads:
dm=podVo=p(t)dV(t)

(A.30)

Here pet) and Po are the densities in the actual and initial states.
Comparing eqns (A.29) and (A.30) yields:
13 =detc=[po/p(t)J2

(A.31)

Using eqn (A. 17) for polar Cayley decomposition and formulae (A.21), we
can also represent the volume-density relations in the equivalent forms:
p(t)/po =(detC)1 /2 = detF= detU=detV

= l/dettfJ = l/detA = l/deti

(A.32)

Here the common formulae related to matrices and their determinants, as


well as the conditions detR=detR*=I, have been taken into account.
When the volume deformations are negligible (in comparison with
shearing ones), eqns (A.31) and (A.32) result in the common condition of
inc om pressi bility:
(A.33)

A1.7 Scalar Functions of Tensors and Their Tensor Derivatives


Besides the tensor functions of the tensor argument considered, there are
also some scalar functions of tensor argument. A good example of such a
function useful in theory of elasticity is the free energy function which
depends on tensor measures of elastic deformations.
Let A be a second-rank symmetric tensor in [R3. If the variations of
components of tensor A in any basis completely predetermine the
variations in a scalar f, it is said that f is a scalar function of tensor A or
f = f(A). If the function f(A) depends only on tensor A and metric tensor
g, it is said that f(A) is an isotropic scalar function of tensor argument.

Appendices

419

These are the functions only of independent say basic invariants, 1 1 , 12


and 13 of tensor A. When dealing with isotropic elastic (usually amorphous) solids, the strain energy function is a typical example of an
isotropic scalar function of tensor argument. When dealing with anisotropic elastic solids (like crystals), the strain energy function is an
anisotropic scalar function of tensor argument. This means that besides
the tensor A and metric tensor g, the scalar function f also depends on
some, usually constant, tensor(s) characterizing the symmetry properties
of unstrained material.
The derivative of a scalar function with respect to tensor argument is a
second-rank symmetric tensor defined as follows:
(A.34)
Due to eqn (A.34), the scalar function f(A) is also called as the potential
function of tensor argument. From now on we will consider as an example
of tensor argument A, the Finger strain tensor c = C- 1 which is positive definite. Then the isotropic scalar function f takes the form
f = f(I 1> 12 , 13 ) where invariants lk are defined by eqns (A.26). In this case,
eqn (A.34) is represented as follows:

df/dc= L Aolk /oc

(A.35)

Simple calculations of the derivatives of basic invariants lk with respect


to tensor c due to eqns (A.26) yield:
(A.36)
The first two relations here are evident. The third one can be easily
derived if we make the trace operation for the Hamilton-Cayley identity
(A.25) and substitute into the result the expression for 12 from eqn (A.26).
In so doing we obtain:
13=1i;6-!Il trc 2 +t trc 3

Differentiating this equality with respect to c gives:


01 3/0C= I 2fJ- 11 c +c 2 == 1 3c- 1

where the last transition in the above chain equality was made due to
Hamilton-Cayley identity (A.25). This result is evident from the viewpoint
of representing 13 through the principal values Ck as shown in eqn (A.27).
Thus eqns (A.35) and (A.36) finally result in:
df/dc = (fl +f212)b-f2c+f313c-l

(A.37)

420

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Notice that eqn (A.37) is a particular case of general representation


(A.28b) for the isotropic tensor function of tensor argument.
A1.8 Strain Rate and Vorticity Tensors
The field of velocity vector v is defined as follows:
(A.38)

v=or(~, t)/otl~
I

This kinematic magnitude is inconvenient to use in the constitutive


equations because its value depends on the choice of the inertial frame of
reference, i.e. it is not invariant relative to Galilean transformations. The
velocity gradient tensor Vv defined by the equality:
(A.39)
is a Galilean invariant. Notice that the tensor bv/br defined by eqn
(A.39) has the covariant components (bv/br)jj = VjVj whereas according
to the common definition of velocity gradient tensor Vv its covariant
components take the form: (Vv)ij = VjVj. This is why we used the transposition symbol t in eqn (A. 39). In a Cartesian coordinate system
(Xl, xz, X3), the matrix of the velocity gradient tensor Vv takes the
form:

OVdOXl
Vv= [

OVZ/OXl

ovdoxz oVz/oxz
OVdOX3

(AAO)

OVZ/OX3

Since the velocity gradient tensor is generally non-symmetrical, it is


usually decomposed in the sum of symmetrical e and antisymmetrical OJ
parts:
Vv=e+OJ
e=![Vv+(Vv)t]
OJ = ![Vv - (Vv)t]

(2ejj =

VjVj

+ VjVj)

(2wjj = VjVj - VjVj)

(AA1)

The strain rate tensor e describes the rate of pure deformations in


macroscopically infinitesimal particles of continuum, while the vorticity
tensor OJ describes the speed of rigid rotations of the continuum particles.
Now we can establish useful relations between the velocity gradient
tensor and the time derivative of strain gradient tensors f/J and F.
Differentiating the second equality in eqns (A.ll) with respect to time,

421

Appendices

under condition
(A.39), yields:

~=

constant and with allowance for eqns (A.38) and


bv = dt/J /dt

b~ =

dt/J/dt t/J -1 br
0

Hereafter in the Eulerian description employed, d/dt ==%t+v V. Thus the


relation searched for and represented in terms of Finger strain tensor, is:
0

(A.42a)
Now differentiating the identity t/J F = 0 with respect to time yields:
0

dt/J/dt t/J - 1 = - F- 1 dF/dt


0

This, upon substituting into eqn (A.42a) gives the equivalent representation of velocity gradient tensor through the Green strain tensor and its
time derivative as follows:
(AA2b)

A1.9 Evolution Equation for Strains

Equation (A.38a) and its transposed equation can be represented in the


form:
dt/J/dt = Vv t t/J,

(A.43)

The second relation (A.43) was obtained from the first one by using the
common tensor operations: (dt/J/dt)t= dt/Jt/dt and (AoB)t=Bt.At. Let us
derive first, the evolution equation for the Finger strain tensor c == C- 1 .
Differentiating the equality c = t/J t/J t (see eqn (A.21a)) with respect to time
gives:
(A.44)
0

With allowance for eqn (A.44), the sum of the first relation in eqns (A.43)
right multiplied by t/J t and the second left multiplied by t/J, yields:
(c==dc/dt)

(AA5a)

Here t is the upper-convected derivative of Finger strain tensor c.


Equation (A.45a), linear in c, demonstrates the conservation of metric
tensor g under operation of the time-convected derivative. Using eqn
(AAO) and the simple operations: et = e, wt = - w, eqn (A.45a) can be also
represented in the equivalent form:
c-cow+woc-eoc-coe==c-coe-eoc=O

(AA5b)

where c is the Jaumann or co-rotational time derivative of tensor c.

422

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Also, representing eqn (AA2) and its transposed equation in the form

dF/dt= -F'(Vv)t,

(A.46)

and using the time differentiation of Green strain tensor, C = Ft. F, as


follows: t:= ftt . F + Ft ft, yields:

C+Vv'C+C'(Vv)t:=C=O

(A.47a)

where C is the lower-convected derivative of Green strain tensor C.


Substituting eqn (AAO) into eqn (AA7a) also gives:

t - C'w+w' C + C e+ e' C:= t


where

C is

+ C'e+e' C=O

(AA7b)

the Jaumann or co-rotational derivative of tensor C.

At.tO Continuity Equation

Evolution equations for strains (AA5) and (AA7) contain all the information of metric properties of continuum including the continuity or mass
conservation equation. It is demonstrated below, for instance, that the
continuity equation is a simple consequence of the evolution equation
(AA5a). For this reason, let us preliminarily establish a useful relation.
Let c = c(t, x), then 13 = 13(C(t, x)). Now differentiating 13 with respect to
time as a complex function, with allowance for the third relations in eqns
(A.26) and (A.36), yields:

d13/dt = tr(c' d1 3/dc) = 13 tr(c' c- 1 )


Using eqn (A.31), we can rewrite the above equation as follows:
(AA8)

where p = p(t, x) is the density of the medium. Now multiplying eqn


(AA5a) scalarly by c- 1 and using eqn (AA8) gives the continuity equation:
dIn p/dt+ V v=o

(A. 49a)

which also is rewritten in more usual form:


op/ot+V'(pv)=O

(AA9b)

For the incompressible case when p = Po = constant (13= 1), the continuity equation takes the common form:
(A. 50)

423

Appendices

A2 A BRIEF INTRODUCTION INTO NON-EQUILIBRIUM


THERMODYNAMICS
A2.1 Conservation Laws: The Local Formulation of The First
Law of Thermodynamics
We will consider the continuum media without internal rotations where
the stress tensor is symmetric: 0' = O't, and the total specific energy per unit
mass, w, is expressed as the sum of internal, u, and kinetic, v 2 /2, energies,
i.e.
PW = pu + pv 2 /2

(A.51)

where v2 = v v and v is the velocity vector. The conservation laws of mass,


momentum and energy, general for these media, have the common local
form:

dp/dt= -pVv,

pdv/dt=VO',

pdw/dt = - V J w

(A.52)

where J w is the total energy flux. Multiplying scalariy the second,


momentum balance eqn (A.52) by velocity vector v gives the equation for
balance of kinetic energy:

pd/dt(v 2 /2) = V (0'. v) - tr(O' e)

(A.53)

Now substituting eqn (A.51) with allowance for eqn (A.53) into the third,
total energy conservation, eqn (A.52) yields the equation for balance of
internal energy:

pdu/dt= - V .Jq + tr(O'e)

(A. 54)

where the thermal flux, J q is related to the total energy flux, J w , as


follows:
(A.55)
Equation (A.55) is, indeed, a definition of energy flux J w Equation (A.54)
is the local formulation of the First Law of Thermodynamics for the
media under study.
A2.2 Local Equilibrium Assumption: Gibbs' Relation. A Local
Formulation of the Second Law of Thermodynamics
When considering the deformation and flow of dissipative media, such
as polymeric fluids, suspension or plastic flows in metals, the very concept of thermodynamics is at least not so clear (and sometimes doubtful) in comparison with the case of thermodynamic equilibrium. The

424

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

fundamental question is, whether and to what extent we can operate in


non-equilibrium with the common thermodynamic functions and the
relations between these which are defined in the state of true equilibrium.
To some extent, this question is clarified by the local equilibrium
assumption and some particular microscopic models of various media (see,
e.g. Ref. 53). According to the hypothesis of local equilibrium, it is usually
assumed that though the dissipative media as a whole are in nonequilibrium, if the deviations from equilibrium are 'not so high',there is still
local thermodynamic equilibrium in infinitesimal macroscopic particles
where all the thermodynamic variables and functions, as well as the
fundamental relation between these, are valid. In application of this
hypothesis to the flow of polymeric liquids, we have considered as local
equilibrium, the state of an elastic rubber-like solids with high elastic
(recovery) strains. At present, when no strict irreversible statistical mechanics have been built up, profound theoretical foundations of this hypothesis
are very slim. Generally, only indirect verifications of the assumption as
compared to theoretical descriptions and observations, can serve to justify
the whole approach. Fortunately, there is a lot of positive evidence for this.
Employing the local equilibrium assumption, we can rewrite all the
differential relations common in equilibrium thermodynamics as respective
time derivatives taken along the path of infinitesimal particles. Consider,
for instance, the dependence u = u(s, Yk) where s is the specific (per unit
mass) entropy and Yk are other state variables. Then the famous Gibbs'
relation
du=Tds+dul s

(T = au/as IYk , du Is =

aU/OYk Is dYk)

(A.56)

where T is temperature. In non-equilibrium, the Gibbs' relation (A.56) is


written as follows:
du/dt = Tds/dt + du/dt Is

(A.57a)

Note that the local entropy s introduced here has a sense only due to the
local equilibrium assumption.
We can also introduce the specific free energy f (per unit mass) by the
common relation:
(A. 58)
for which
df=du-Tds-sdT= -sdT+dul s
df= af/oTI Yk dT + af/OYk iT dYk

(A.59)

425

Appendices

Comparing these two relations with allowance for eqn (A.56) yields:
dflT=dul s

In non-equilibrium, we can rewrite eqn (A.59) in the dual Gibbs' form:


df/dt= -sdT/dt + df/dt IT

(A.57b)

Now we can write the Second Law of Thermodynamics in the local form:
pds/dt= -VJs+P.,

(A.61)

where J s is the entropy flux and P s is the entropy production related to


dissipative processes in the medium. According to the Second Law, P s >0
for all non-equilibrium processes and vanishes in equilibrium.
A2.3 Expressions for Entropy Production, Entropy Flux and
Heat Capacity
Substituting the Gibbs' relation (A.57a) into eqn (A.54) with allowance for
eqn (A.60) yields:
p ds/dt= - V (Jq/T)+ T- 1 [

T-1Jq. VT +tr(o- e)- p df/dtI T ] (A.62)

Equation (A.62) has the same form as eqn (A.61) proposed for the entropy
balance. Comparing the respective terms gives:
TP s = - T-1Jq. VT+tr(o-e)-pdf/dtI T
(pdf/dtIT=

(A.63)

L Of/OYkITYk)
k

When temperature inhomogeneities are negligible, we call the magnitude


TPsIT=D the dissipative function. The apparent uncertainty in the decomposition of the right-hand side of eqn (A.62) into 'origin and divergent'
parts, is eliminated on the reasons that entropy production Ps vanishes in
equilibrium and is invariant under Galilean transformations, i.e. is
invariant relative to choice of inertial frame of reference.
The heat capacity CYk at constant values of state variables Yk is defined as
in equilibrium by the common relation:
(A.64)
A2.4 Generalized Thermodynamic Forces and Fluxes
According to eqn (A.63), the entropy production Ps is represented as a
characteristic bilinear form LnXnYn where the magnitudes Xn are called
generalized thermodynamic forces and the magnitudes Yn conjugated to

426

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

them are called generalized thermodynamic fluxes. Usually (though not


always), thermodynamic fluxes have the sense of generalized velocities and
therefore they change the sign under transformation t-+ - t. For instance,
the magnitudes: e, }\ and, for some reasons, JqjT, independently from their
tensor dimensionality, are usually named as thermodynamic fluxes.
Conjugated to these magnitudes: (1, o!joYklT and VT are usually named as
thermodynamic fluxes. The quasilinear relations between the thermodynamic forces and fluxes of the same tensor dimensionality (which is
due to the Curie principle)53 through kinetic coefficients which are generally
tensor dependent on the state variables, are called the constitutive equations.
In the non-equilibrium thermodynamic approach considered, these constitutive equations are also subjected to an additional constraint related to so
called Onsager's symmetry principle. 53 This quasilinear approach is widely
used in the main text to obtain the various constitutive equations.

A3 COMPONENT-WISE EXPRESSIONS FOR BASIC


EQUATIONS
A3.1 The Equations of Momentum Balance and Continuity
for Incompressible Media

pdvjdt=V(1+pg,

Vv=o

(A.65)

where g is the gravity acceleration. The stress tensor is represented as


follows:
(A. 66)
where p is a scalar pressure. The 'extra-stress' tensor 'C is related to
rheological properties of incompressible media, while the pressure p is
usually determined from boundary conditions. Equations (A.65), with
allowance for eqn (A.66), are represented below in rectangular, cylindrical
and spherical coordinate systems.

Rectangular coordinates (x, y, z):


OV x
oVx
oV x
OV x) _ op or xx or yX or zx
p ( ar+ vx ox +Vy oY +vza; - - ox + ox + oY + oz +pgx
OVy
OVy
oVy
Ovy) _ op or XY oryy or zy
p ( ar+ vx ox +Vy oY +vza; - - oY + ox + oY + oz +pgy

427

Appendices

(A.67)

Cylindrical coordinates (r, cp, z):


Ovr
oVr vq> oVr v~
OVr)
p ( -+V -+- ---+V ot
r or
r ocp r
z OZ
op
or

1 0

1 07: r 7:pp
r or
r ocp
r
OV",
oVq> v",ovq> vrvq>
OVq
p ( -+vr-+--+-+vzot
or r ocp
r
OZ

07: rz
OZ

= - -+- -(rLrr)+- - " ' - - +-+pgr

1 op 1 0 2
07:'1''1' 07:q>z
-+- -(r 7: r )+-1 -+-+pg
r ocp r2 or
'" r ocp
OZ
'I'

= --

OVz
OVz Vq> OVz
OVz)
P( -;-+Vr-;-+--;-+Vz-;ut
ur r ucp
uZ
Op
OZ

1 0
r or

1 07:q>z
r ocp

07: zz
OZ

= - -+- -(rLrz)+- -+-+pgz

1 0
1 OVq> OV z
- -(rVr)+- -+-=0
r or
r ocp OZ

Spherical coordinates (r, cp, e):


Ovr
OVr V9 OVr
Vq> OVr V~ + V~)
p ( at + V'a;:- + -;:- Oe + r sin e ocp - - r op
ur

1 0

= --;-+2 -;-(r 7: rr )+-'-e :1e(7:r9 sme)

r ur

r sm

1
07:r", 7:'1''1'+'99
+-'-e --;-+ pgr
r sm ucp
r

OV9
at

OV9
r ar

V9 OVq>
v ae

Vq> OV,9 VrV9


r sin e acp
r

v; cot e)
r

p ( -+V - + - - + - - - + - - - ' - - -

lap 1 a 2
1
a
.
= ----;-e+
2 -;-(r 7:r9 )+-'-e :1e(7:99 sme)
ru
r ur
' r sm u

(A.68)

428

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

1 0'tp8 'r9 cot 0


+ r sin 0 o<p +-;---r- 'tptp + pg9
OVtp
ot

oVtp
or

V9 oVtp
Vtp oVtp
r 00 r sm 0 o<p

vrvtp
r

(A.69)

vtpva
r

p ( -+vr - + - - + - . - - + - + - cot

0)

= __1_ op +~ !(r 2, )+~ 0'tp9 +_1_ o'tptp


r sin 0 o<p

r2 or

r",

r 00

r sin 0 o<p

'r",
2 cot 0
+-;:-+--r-'tp8 + pgtp

1 0 2
1
O.
1 oVtp
-;-:(r vr)+-,-O "'0(V9 sm 0)+-'-0 -;-=0
r ur
r sm u
r sm u<p

"2

A3.2 Equation for Temperature Variations

This equation takes the invariant form (see Chapter 9):


pcdTjdt=RAT + \{'

(A.70)

where, under an assumption of entropy elasticity, the heat generation


term \{' is written as follows:
(A.71)
Here c is the heat capacity and R is the thermal conductivity. Below, the
temperature eqn (A.70) with an allowance for eqn (A.71) is represented in
rectangular, cylindrical and spherical coordinate systems.
Rectangular coordinates (x, y, z):

(A.72)

Cylindrical coordinates (r, <p, z):

429

Appendices

(A.73)

Spherical coordinates (r, cp, 0):

_[ 1

=K

a( ar
aT) + r2 sinO
1
a(. 0 aT)
00
00

r2 or r

SID

1
OV'" Vr Va
)
+u"'''' ( -r 'SID
- n -;-+-+-cotO
u ucp
r r

+ura(ova+! ovr_Va)+ur (ov",+_~_ ovr_v",)


or

r 00

'" or

ov'"+ - .1 - ---cotO
OVa v'"
)
+u a( -1 '" r 00

SID

0 ocp

SID 0

ocp

(A.74)

A3.3 Component-wise Expressions for the Velocity Gradient Tensor


and the Upper-Convected Time Derivative of Recoverable
Strain Tensor
The upper-convected time derivative of recoverable strain tensor c is
represented as follows:

c=oc/ot+v Vc,

(A.75)

430

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

where Vv is the gradient velocity tensor. Below these tensors are represented in component-wise form in rectangular, cylindrical and spherical
coordinate systems.
Rectangular coordinates (x, y, z):

(Vv)YX =

ovx
oy'
ovx

(Vv)zx=az-'

(Vv)yy=

oVy
oy'
ovy

(Vv)ZY=az-'

(Vv)yz=

ov z
oy'
oV z

(Vv)zz=az-

(i,j=x, y,z)

(A.76)

(A. 77)

(A. 78)

431

Appendices

Cylindrical Coordinates (r, <p, z):


OV r
(Vv)rr=-;-,
,
ur

(A.79)

oC

v
rr
(c)rr=~+(v

ut

oV r
1 oV r
oV r
V)c rr - 2c rr -;-- 2crtp - -;--2crz -;ur
r u<p
uZ

(A.80)

(1

v
oCrtp
(OVtp- Vtp) -C r - oVtp+ Vr +OVr)
(c)r
=-+(vV)c
rtp -Crr tp ot
or r
tp r o<p r or

where
(A.81)

432

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Spherical Coordinates (r, ep, 8):

1
oV r v",
(Vv)",r=-'-ll -;--r sm u uep r

1
OV'" Vr Vs
(Vv)"''''=-'-ll -;- + -+- cot 8,
r sm u uep
r r

(Vv)

'"

1
OV.9
r sin 8 oep

v",
cot 8
r

s=-- - - -

1 oVs Vr
(Vv)ss=- - + r 08
r

(A.82)

v
oCrr
oV r
1 oV r
(c)rr =---;-+ (v V)c rr - 2c rr --;-- 2c r", -'-ll -;ut
ur
r sm u uep

OV'"
-2c ( -1- + -Vr+Va
- cot 8) -2c s-1 (OV",
- - v cot8 )
"'''' r sin 8 oep r r
'" r 08
'"
v
OC.99
vs)
(chs=-+(v
V)css - 2car (ovs
--;--ot
ur r

(A.83)

r
OV+
ov",
- 1. - + -Vr+Va
- cot f} )
('" or r sm 8 oep r r

-C r

-c

1
oV r
1 oV r
----cs-"'''' r sin f} oep
'" r 08

1 (OV",
- - v cot 8)
r 08
'"

-C r9-

Appendices

433

v"')
1 (av",
)
( av",
or r -C99-r --v
ae '" cot e
v
aCr9
(av9 v",) - Cr'" -.-e
1
aV9
(C)r9
= -;- + (v V)Cr
Crr --;- - vt
vr r
r SIll -;vcp
1_ aVr",
_C9(avr+~ aV9+~)_c
or r ae r '" r sin e acp
-C9S-1r -aVr
ae
-C r 9 - - -

9-

9_

where

a Vs a + v",- - a
or r ae r sin e acp

(VV)=vr -+-

(A.84)

References

1. Vinogradov, G.V. and Malkin, A. Ya., Rheology of Polymers, Mir Publishers, Moscow, 1980.
2. Pearson, IRA., Mechanics of Polymer Processing, Elsevier, New York,
1985.
3. Janeschitz-Kriegl, H., Polymer Melt Rheology and Flow Birefringence,
Springer, New York, 1983.
4. Tanner, RI., Engineering Rheology, Oxford Press, Oxford, 1985.
5. Bird, RB., Armstrong, RC. and Hassager, 0., Dynamics of Polymeric
Liquids: Vol. I, Fluid Mechanics; Vol. II, Kinetic Theory, 2nd edn, John
Wiley, New York, 1987.
6. Larson, RG., Constitutive Equations for Polymer Melts and Solutions,
Butterworth, Boston, 1989.
7. White, J.L., Principles of Polymer Engineering Rheology, John Wiley, New
York, 1990.
8. Leonov, A.I., Nonequilibrium thermodynamics and rheology of viscoelastic
polymer media. Rheol. Acta, 15 (1976) 85.
9. Leonov, A. I., On a class of constitutive equations for viscoelastic liquids. J.
Non-Newt. Fluid Mech., 25 (1987) 1.
10. Godunov, S.K., Elements of Continuum Mechanics, Nauka, Moscow, 1978
(in Russian).
11. Giesekus, H., Die elastizitat von flussigkeiten. Rheo!. Acta, 5 (1966) 29.
12. Giesekus, H., A unified approach to a variety of constitutive models for
polymer fluids based on the concept of configuration-dependent molecular
mobility. Rheol. Acta, 21 (1982) 366.
13. Isayev, A.I. and Upadhyay, R.K., Two-dimensional viscoelastic flows: experimentation and modeling. J. Non-Newt. Fluid Mech., 19 (1985) 135.

References

435

14. Aybinder, S.B., Tyunina, E.L. and Tsirule, K.I., Properties of Polymers in
Various States, Khimiya, Moscow, 1981 (in Russian).
15. Isayev, AI., (Ed.) Injection and Compression Molding Fundamentals, Marcel
Dekker, New York, 1987.
16. Treloar, L.R.G., The Physics of Rubber Elasticity, 3rd edn, Clarendon Press,
Oxford, 1975.
17. Malkin, A.Ya. and Leonov, A.I., On the instability criteria for shear
deformations of viscoelastic polymer media. Doklady AN SSSR, 151 (1963)
380 (in Russian).
18. Vinogradov, G.V., Malkin, A.Ya. and Leonov, AI., Conditions of unstable
flow of elastoviscous polymeric media. Kolloid Z. und Z. fur Polymere, 191
(1963) 25.
19. Reiner, M., The Deborah number. Physics Today, 17 (1964) 62.
20. Williams, M.L., Landel, R.F. and Ferry, J.D., The temperature dependence
of relaxation mechanism in amorphous polymers and other glass-forming
liquids. J. Am. Chem. Soc., 77 (1955) 3701.
21. Weissenberg, K., A continuum theory of rheological phenomena. Nature,
159 (1947) 310.
22. Kuvshinsky, Ye.V., The study of outflow of high polymer solutions (mechanics of elastic and high elastic viscous media). Doctoral Dissertation,
Phys.-Techn. Institute of Acad. Sci. USSR, Leningrad, 1950 (in Russian).
23. Gorodtsov, V.A. and Leonov, A.I., On kinematics, nonequilibrium thermodynamics and rheological relations in nonlinear theory of viscoelasticity.
PMM: J. Appl. Math. Mech., 32(1) (1968) 70 (in Russian).
24. Leonov, AI., On the equivalence of simplest nonlinear rheological equations for viscoelastic polymer media. Rheo/. Acta, 21 (1982) 683.
25. Leonov, A.I., Quasilinear thermodynamics and rheology in nonlinear theory
of viscoelasticity with the use of recovery strain as an internal parameter.
Ann. N.- Y. Acad. Sci., 410 (1983) 23.
26. Buyevich, Yu. A, On the kinematics for elastoviscous media with finite
strains. PMM: J. Appl. Math. Mech., 32(2) (1968) 217 (in Russian).
27. Dashner, P.A. and Van Arsdale, W.E., A phenomenological theory for
elastic liquids. J. Non-Newt. Fluid Mech., 8 (1981) 59, 69.
28. Stickforth, J., The rational mechanics and thermodynamics of polymeric
fluids based upon the concept of variable relaxed state. Rheol. Acta, 25
(1986) 447.
29. Biot, M.A., Theory of stress-strain relations in anisotropic viscoelasticity
and relaxation phenomena. J. Appl. Phys., 25(11) (1954) 1385.
30. Biot, M.A., Variational principles in irreversible thermodynamics with
application to viscoelasticity. Phys. Rev., 97(6) (1955).
31. Biot, M.A., Mechanics of Incremental Deformations, John Wiley, New York,
1965.
32. Kluitenberg, G.A. Thermodynamic theory of elasticity. Physica, 28 (1962)
217.
33. Kluitenberg, G.A., A note of the thermodynamics of Maxwell bodies, Kelvin
bodies (Voigt bodies) and fluids. Physica, 28 (1962) 561.
34. Kluitenberg, G.A., On the rheology and thermodynamics of irreversible
processes. Physica, 28 (1962) 1173.

436

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

35. Kluitenberg, G.A., On the thermodynamics of viscosity and plasticity.


Physica, 29 (1963) 633.
36. Kluitenberg, G.A., A unified thermodynamic theory for large deformations
in elastic media and in Kelvin (Voigt) media, and for viscous fluid flow.
Physica, 30 (1964) 1945.
37. Kluitenberg, G.A., Application of the thermodynamics of irreversible processes to continuum mechanics. In Non-Equilibrium Thermodynamics, Variational Technique and Stability, The University of Chicago Press, Chicago,
IL, 1966, p. 91.
38. Leonov, A.I. and Prokunin, A.N., The improved simple version of a
nonlinear theory of elastoviscous polymer media. Rheol. Acta, 19 (1980)
393.
39. Prokunin, A.N., On nonlinear Maxwell-like constitutive equations for the
description of flows of polymer fluids. PMM: J. Appl. Math. Mech., 48(6)
(1984) 957 (in Russian).
40. Prokunin, A.N., On the description of viscoelastic flows of polymer fluids.
Rheol. Acta, 23 (1989) 38.
41. Leonov, A.I., Analysis of simple constitutive equations for viscoelastic
liquids. J. Non-Newt. Fluid Mech., 42 (1992) 323.
42. Oldroyd, J.G., The search for rheological equations of state consistent with
thermodynamics. In Proc. VII Intern, Congr. Rheology, Tages-Anzeiger,
Zurich, Switzerland, 1976, p. 102.
43. Eckart, c., The thermodynamics of irreversible processes. IV. The theory of
elasticity and anelasticity. Phys. Rev., 73(2) (1948) 373.
44. Truesdell, c., A First Course in Rational Continuum Mechanics, Vol. 1,
General Concepts, Academic Press, New York, 1977.
45. Truesdell, c., Rational Thermodynamics, McGraw-Hill, New York, 1969.
46. Kondo, K. (Ed.), Non Riemannian Geometry of Imperfect Crystals from a
Macroscopic Viewpoint. Memoirs of the Unifying Study of the Basic Problems
in Engineering Sciences by Means of Geometry, Vol. 1, Gakujutsu Bunken
Fukuyu-Kai, Tokyo, 1955.
47. Bilby, B.A., Bullogh, R. and Smith, E., Continuous distribution of dislocations; a new application of the methods of non-Riemannian geometry.
Proc. Roy. Soc. London, A231 (1955) 263.
48. Kroner, E. and Seeger, A., Dislocations and stresses in nonlinear theory of
elasticity. Arch. Rat. Mech. Anal., 3 (1959) 97.
49. Kroner, E., General continuum theory of dislocations and stresses. Arch.
Rat. Mech. Anal., 4 (1960) 273.
50. Besseling, J.F., A thermodynamic approach to rheology. In A Thermodynamic Approach to Rheology. Irreversible Aspects of Continuum Mechanics and Transfer of Physical Characteristics in Moving Fluids (Ed. H.
Park us and L.I. Sedov), IUTAM Symposia, Springer, Vienna, 1968, p. 16.
51. Lee, E.H. and Liu, D.T. Finite strain elastic-plastic theory. In Irreversible
Aspects of Continuum Mechanics and Transfer of Physical Characteristics in
Moving Fluids, (H. Park us and L.I. Sedov), IUTAM, Symposia, Springer,
Vienna, 1968, p. 213.
52. Murnaghan, F.D., Finite deformations of elastic solids. Amer. J. Math., 59
(1937) 235.

References

437

53. De Groot, S.R and Mazur, P., Non-Equilibrium Thermodynamics, NorthHolland, Amsterdam, 1962.
54. Hellwege, KH., Henning, I. and Knappe, W., The anisotropy of thermal
expansion and thermal conductivity in amorphous high polymers that are
stretched along one axis. Kolloid-Z. Polymere, 188 (1963) 121.
55. Valanis, KC. and Landell, RF., The strain-energy function of a hyperelastic
material in terms of the extension ratios. J. App/. Phys., 38 (1967) 2997.
56. Ogden, RW., Large deformation isotropic elasticity - on the correlation of
theory and experiment for incompressible rubberlike solids. Proc. Roc. Soc,
A326 (1972) 565.
57. Blatz, P.l, Sharda, S.G. and Tschoegl, N.W., Strain energy function for
rubberlike materials based on a generalized measure of strain. Trans. Soc.
Rheol., 18 (1974) 94.
58. Palmov, V.A., Rheological Models in nonlinear mechanics of deformed
media. Advances in MechaniCS, 3(3) (1980) 75 (in Russian).
59. Ziegler, H., Some extremum principles in irreversible thermodynamics with
application to continuum mechanics. Progress in Solid Mechanics, 4 (1963) 91.
60. Larson, RG., Elongational-flow predictions of the Leonov constitutive
equation. Rheol. Acta, 22(5) (1983) 435.
61. Glasston, S., Laidler, KJ. and Eyring, H., The Theory of Rate Processes,
McGraw Hill, New York, 1941.
62. Frenkel, Ya.N., Kinetic Theory of Liquids, Nauka, Leningrad, 1975 (in
Russian).
63. Gross, B., Mathematical Structure of the Theories of Viscoelasticity, Hermann, Paris, 1953.
64. Bland, D., The Theory of Linear Viscoelasticity, Pergamon Press, Oxford,
1960.
65. Leonov, A.I., Lipkina, E.Kh., Paskhin, E.D. and Prokunin, A.N., Theoretical and experimental investigations of shearing in elastic polymer liquids.
Rheol. Acta, 15 (1976) 411.
66. Ward, I.M., Mechanical Properties of Solid Polymers, 2nd edn, John Wiley,
New York, 1983.
67. Bartenev, G.M., Effects of the shearing stress on the viscosity of linear
polymers. Vysokomolekularnye Soyedinenya, 6 (1964) 2155 (in Russian).
68. Frohlich, H. and Sack, R, Theory of rheological properties of dispersions.
Proc. Roy. Soc., A185 (1946) 415.
69. Oldroyd, J.G., On the formulation of rheological equations of state. Proc.
Roy. Soc., A200 (1950) 1063.
70. Oldroyd, J.G., Non-Newtonian effects in steady motion of some idealized
elastico-viscous liquids. Proc. Roy. Soc., A245 (1958) 278.
71. Rivlin, R.S., Research Frontiers in Fluid Dynamics, Wiley Interscience, New
York, 1965, p. 144.
72. Lodge, A.S., Elastic Liquids, Academic Press, London, 1964.
73. Fredrickson, A.G., Principles and Applications of Rheology, Prentice-Hall,
New York, 1964.
74. Gordon, RJ. and Schowalter, W.R, Anisotropic fluid theory: a different
approach to the dumbbell theory of dilute polymer solutions. Trans. Soc.
Rheol., 16 (1972) 79.

438

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

75. Johnson, Jr., M.W., and Segalman, D., A model for viscoelastic fluid
behavior which allows non-affine deformations. J. Non-Newt. Fluid Mech., 2
(1977) 255.
76. Phan Thien, N. and Tanner, R.I., A new constitutive equation derived from
network theory. J. Non-Newt. Fluid Mech., 2 (1977) 353.
77. Phan Thien, N., A nonlinear network viscoelastic model. J. Rheol, 22 (1978)
259.
78. White, IL. and Metzner, A.B., Development of constitutive equations for
polymeric melts and solutions. J. Appl. Polym. Sci., 7 (1963) 1867.
79. Marucci, G., Molecular modeling of flows of concentrated polymers. In
Advances in Transport Processes, Vol. 5 (ed. A.S. Mujamdar and R.A.
Mashelkar), John Wiley, New York, 1984.
80. Larson, R.G., A constitutive equation for polymer melts based on partially
extending strand convection. J. Rheol, 28 (1984) 545.
81. Grmela, M., Bracket formulation of diffusion--convection equations.
Physica, 2tD (1986) 179.
82. Grmela, M., Hamilton dynamics of incompressible elastic liquids. Phys.
Lett., A130 (1988) 81.
83. Beris, A.N. and Edwards, B.J., Poisson bracket formulation of incompressible flow equations in continuum mechanics. J. Rheol., 34 (1990) 55.
84. Beris, A.N. and Edwards, B.1., Poisson bracket formulation of viscoelastic
flow equations of differential type: a unified approach. J. Rheol., 34 (1990)
503.
85. Morrison, P.J. and Greene, J.M., Noncanonical Hamilton density formulation of hydrodynamics and ideal magnetohydrodynamics. Phys. Rev. Lett,
45 (1980) 790.
86. Dzyaloshinskii, I.E. and Volovick, G.E., Poisson brackets in condensed
matter physics. Ann. Phys., 125 (1980) 67.
87. Rivlin, R.S. and Sawyers, K.N., Nonlinear continuous mechanics of viscoelastic liquids. Ann. Rev. Fluid Mech., 3 (1971) 117.
88. Goddard, J.D. and Miller, C, An inverse for the Juamann derivative and
some applications to the rheology of viscoelastic fluids. Rheol. Acta, 5 (1966)
177.
89. Kaye, A., Non-Newtonian flow in incompressible fluids, Part I: A general
rheological equation of state; Part II: Some problems in steady flow. College
of Aeronautics, Crawford, UK, Note no. 134, 1962.
90. Berstein, B., Kersley, E.A. and Zapas, L.1., A study of stress relaxation with
finite strain. Trans. Soc. Rheol., 7 (1963) 391.
91. Wagner, M.H. Raible, T. and Meissner, J., Tensile stress overshoot in
uniaxial extension of LDPE melt. Rheol. Acta, 18 (1979) 427.
92. Wagner, M.H. and Dermamels, A., A constitutive analysis of extensional
flows of polyisobutylene. J. Rheol., 34 (1990) 943.
93. Truesdell, C and Noll, W., The Non-Linear Field Theories of Mechanics,
Springer Verlag, Berlin, 1965
94. Astarita, G. and Marucci, G., Principles of Non-Newtonian Fluid Mechanics,
McGraw-Hill, New York, 1974.
95. Zaremba, S., On a perfect form of the theory of relaxation. Bull. Int. Acad.
(Cracovie), (1903) 594 (in French).

References

439

96. Jaumann, G., A closed set of physical and chemical laws formulated in
differential form. Sitzungsberichte Akad. Wiss. Wien, 120 (1911) 385 (in
German).
97. Joseph, D.D., Fluid Mechanics of Viscoelastic Liquids, Springer Verlag, New
York,1990.
98. Volterra, V. and Peres, J., Theorie genera Ie des fonctionels, GauthierVillars, Paris, 1936, p. 61.
99. Coleman, B.D., Thermodynamics of materials with memory. Arch. Rat.
Mech. Anal., 17 (1964) 1.
100. Coleman, B.D., On thermodynamics, strain impulses and viscoelasticity.
Arch. Rat. Mech. Anal., 17 (1964) 230.
101. Coleman, B.D., Markovitz, H. and Noll, W., Viscometric Flows of Nonnewtonian Liquids, Springer Verlag, Berlin, 1966.
102. Coleman, B.D. and NolJ. W., On approximation theorem for functionals,
with applications in continuum mechanics. Arch. Rat. Mech. Anal., 6 (1960)
355.
103. Volkenstein, M.V., Configurational Statistics of Polymeric Chains, Interscience, New York, 1963.
104. Flory, P.I, Principles of Polymer Chemistry, CornelJ University Press,
Ithaka, 1953, p. 432.
105. Altenberg, A.R. and Dahler, J.S., Statistical mechanics of rubber elasticity. J.
Chem. Phys., 92 (1990) 3100.
106. Kargin, V.A. and Slonimsky, G.L., On the deformation of amorphous-liquid
polymers. Doklady AN SSSR 62 (1948) 239 (in Russian).
107. Kargin, V.A. and Slonimsky, G.L., On the determination of polymer
molecular weight using mechanical properties of these. J. Phys. Chem., 28
(1949) 563 (in Russian).
108. Gotlib, Yu.Ya. and Volkenstein, M.V., The theory of relaxation spectrum of
polymeric chain. J. Techn. Phys., 23 (1953) 1936 (in Russian).
109. Kirkwood, J.G. and Riseman, J., The intrinsic viscosities and diffusion
constants of flexible macromolecules in solution. J. Chem. Phys., 16 (1948)
565.
110. Rouse Jr., P.E., A theory of the linear viscoelastic properties of dilute
solutions of coiling polymers. J. Chem. Phys., 21 (1953) 1272.
111. Zimm, B., Dynamics of polymer molecules in dilute solution: viscoelasticity,
flow birefringence and dielectric loss. J. Chem. Phys., 24 (1956) 269.
112. Cerf, R., La macromolecule en chaine dans un champ hydrodynamique.
Theorie generale. Proprietes dynamo-optiques. J. Polymer Sci, 23 (1957)
125.
113. Chandrasekhar, S., Stochastic problems in physics and astronomy. Rev.
Mod. Phys., 15 (1943) 1.
114. Baldwin, P.R. and Helfand, E., Dilute polymer solution in steady shear flow:
Non-Newtonian stress. Phys. Rev., A41 (1990) 6727.
115. de Gennes, P.-G., Scaling Concepts in Polymer Physics, CornelJ University
Press, Ithaka, 1979.
116. Volkov, V.S., The present-day aspect of the structural approach to the
theory of viscoelasticity of linear polymers. Int. J. Polym. Mater., 9 (1982)
115.

440

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

117. Volkov, V.S., Slow flow of a sphere in an anisotropic viscoelastic liquid.


PPM: Appl. Math. Mech., 40 (1982) 248 (in Russian).
118. Green, M.S. and Tobolsky, A.V., A new approach to the theory of relaxing
polymeric media. J. Chern. Phys., 14 (1946) 80.
119. Lodge, A.S., A network theory of flow birefringence and stress in concentrated polymer solutions. Trans. Faraday Soc., 52 (1956) 120.
120. Lodge, A.S., Constitutive equations from molecular network theories for
polymer solutions. Rheol. Acta, 7 (1968) 379.
121. Yamamoto, M., The viscoelastic properties of network structure. I. General
formalism. J. Phys. Soc. Japan, 11 (1956) 413.
122. Yamamoto, M., The viscoelastic properties of network structure. II. Structural viscosity. J. Phys. Soc. Japan, 12 (1957) 1148.
123. Yamamoto, M., The viscoelastic properties of network structure. III. Normal stress effect (Weissenberg effect). J. Phys. Soc. Japan, 13 (1958) 1200.
124. Doi, M. and Edwards, S.F., Dynamics of concentrated polymer systems.
Part 1 - Brownian motion in the equilibrium state. J. Chern. Soc., Faraday
Trans. 11, 74 (1978) 1789.
125. Doi, M. and Edwards, S.F., Dynamics of concentrated polymer systems.
Part 2 - Molecular motion under flow. J. Chern. Soc., Faraday Trans. 11, 74
(1978) 1802.
126. Doi, M. and Edwards, S.F., Dynamics of concentrated polymer systems.
Part 3 - The constitutive equation. J. Chern. Soc., Faraday Trans. 11, 74
(1978) 1818; 75 (1979) 32.
127. de Gennes, P.G., Reptation of a polymer chain in the presence of fixed
obstacle. J. Chern. Phys., 55 (1971) 572.
128. Viory, J.L., Monnerie, L. and Tassin, J.F., Tube relaxation: a necessary
concept in the dynamics of strained polymers. J. Polyrn. Sci., Polyrn. Phys.
Ed., 21 (1983) 2427.
129. McLeish, T.C.B. and Ball, R.C., A molecular approach to the spurt effect in
polymer melt flow. J. Polyrn. Sci., Part B. Polymer Physics, 24 (1986) 1735.
130. Yarin, A.L., Strong flows of polymeric liquids. Part 1. Rheological behavior.
J. Non-Newt. Fluid Mech., 37 (1990) 113.
131. Curtiss, C.F. and Bird, R.B., A kinetic theory for polymer melts. I. The
equation for the single-link orientational distribution function. J. Chern.
Phys., 74 (1981) 2016.
132. Curtiss, C.F. and Bird, R.B., A kinetic theory for polymer melts. II. The stress
tensor and rheological equation of state. J. Chern. Phys., 74 (1981) 2026.
133. Pokrovsky, V.N. and Volkov, B.S., On the theory of slow relaxation
processes in linear polymers. Visokornolekularniye Soyedinenia, A20 (1978)
253 (in Russian).
134. Volkov, V.S. and Vinogradov, G.V., Theory of dilute polymer solution for a
viscoelastic fluid with a single relaxation time. J. Non-Newt. Fluid Mech., 15
(1984) 23.
135. Volkov, V.S. and Vinogradov, G.v., Relaxation interactions and viscoelasticity of polymer melts. Part I: Model development. J. Non-Newt. Fluid
Mech. 18 (1986) 163.
136. Volkov, V.S. and Vinogradov, G.V., Theory of relaxation interaction in
linear polymers. J. Polyrn. Sci., Part B, Polymer Physics, 24 (1986) 2073.

References

441

137. Volkov, V.S. and Vinogradov, G.V., Molecular theories of nonlinear viscoelasticity of polymers. Rheol. Acta, 26 (1988) 231.
138. Dupret, F. and Marchal, J.M., Loss of evolution in the flow of viscoelastic
liquids. J. Non-Newt. Fluid Mech., 20 (1986) 143.
139. Rutkevich, I.M., Some general properties of the equations of viscoelastic
fluid dynamics. Prikladnaya Matematika i Mekhanika (PMM), 33 (1969) 30.
140. Rutkevich, I.M., Steady flow of a viscoelastic fluid in a channel with
permeable walls. Prikladnaya Matematika i Mekhanika (PMM), 33 (1969)
573.
141. Rutkevich, I.M., The propagation of small perturbations in a viscoelastic
fluid. Prikladnaya Matematika i Mekhanika (PMM) 34 (1970) 35.
142. Rutkevich, I.M., On the thermodynamic interpretation of the evolutionary
conditions of the equations of the mechanics of finitely deformable viscoelastic media of Maxwell-type. Prikladnaya Matematika i Mekhanika
(PMM), 36 (1972) 283.
143. Richtmeyer, R.D. and Morton, K., Difference Methods for Initial Value
Problems, Interscience, New York, 1962.
144. Whitham, G.B., Linear and Nonlinear Waves, John Wiley, New York, 1974,
Chapter 10.
145. Godunov, S.K., Thermodynamics of gases and differential equations. Uspekhi Matematicheskikh Nauk, 5 (1959) 97 (in Russian).
146. Godunov, S.K., Equations of Mathematical Physics, Nauka, Moscow, 1971
(in Russian).
147. Truesdell, c., The mechanical foundations of elasticity and fluid dynamics,
Chapter IV. Elasticity. J. Rat. Mech. Anal., 1 (1952) 173.
148. Green, A.E. and Adkins, lE., Large Elastic Deformations and Non-Linear
Continuum Mechanics, Clarendon Press, Oxford, 1960, Chapter IX.
149. Beatty, M.L., Some static and dynamic implications of the general theory of
elastic stability. Arch. Rat. Mech. Anal., 19 (1965) 167.
150. Beckenbach, E.F. and Bellman, R., Inequalities, Springer Verlag, Berlin,
1961, Chapter 2.
151. Kwon, Y. and Leonov, A.I., On Hadamard-type stability of single integral
constitutive equations for viscoelastic liquids. J. Non-Newt. Fluid Mech. 47
(1993) 77.
152. Hulsen, M.A., A sufficient condition for a positive definite configuration
tensor in differential models. J. Non-Newt. Fluid Mech., 38 (1990) 93.
153. Joseph, D.D. and Saut, J.c., Change of type and loss of evolution in the
flows of viscoelastic fluids. J. Non-Newt. Fluid Mech., 20 (1986) 117.
154. Kolkka, R.W., Malkus, D.S., Hansen, M.G., lerley, G.R. and Worthing,
R.A., Spurt phenomena of the Johnson-Segalman fluid and related models.
J. Non-Newt. Fluid Mech., 29 (1988) 303.
155. Malkus, D.S., Nohel, J.A. and Plohr, B.J., Dynamics of shear flow of a
non-Newtonian fluid. J. Comput. Phys., 87 (1989) 464.
156. Belkin, I.M., Vinogradov, G.V. and Leonov, A.I., Rotation Devices.
Measurements of Viscosity and Physico-Mechanical Characteristics of Materials, Mashinostroyeniye, Moscow, 1968 (in Russian).
157. Middleman, S., The Flow of High Polymers. Continuum and Molecular
Theories, Wiley, New York, 1968.

442

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

158. Prokunin, A.N., Petrosyan, A.Z. and Fridman, M.L., On the dependences of
viscosity of plastisoles on time and temperature. Inzenerno-Fizicheskii
Journal, 51 (1986) 609 (in Russian).
159. Marsh, B.D. and Pearson, J.R., The measurement of normal stress differences using a cone-and-plate total thrust apparatus. Rheol. Acta, 7 (1968)
327.
160. Petersen, J.F., Rautenbach, R. and Schummer, P., Zur Bestimmung der
Normalspannungs-Funktionen von Hochpolymeren mittels der KegelPlatte-Abstand-Anordnung (KPA). Rheo!. Acta, 14 (1975) 968.
161. Yakobson, E.E. and Faytelson, L.A., Stressed-deformed state of high molecular liquids at steady state shear flow and the stored energy. Mechanics of
Composite Materials, 2 (1985) 328 (in Russian).
162. Ferry, J.D., Viscoelastic Properties of Polymers, Wiley, New York, 1970.
163. Maxwell, B. and Chartoff, R.P., Studies of a polymer melt in an orthogonal
rheometer. Trans. Soc. Rheol., 9 (1965) 41.
164. Blyler, L.L. and Kurtz, S.J., Analysis of the Maxwell orthogonal rheometer.
J. Appl. Polym. Sci., 11 (1967) 127.
165. Zabugina, M., Experimental investigation of high elasticity in flowing
polymers. PhD Thesis, Institute of Petrochemical Synthesis, USSR Academy
of Science, Moscow, 1972 (in Russian).
166. Lee, K.H., Jones, L.G., Pandalai, K. and Brodkey, R.S., Modifications of an
R-16 Weissenberg Rheogoniometer. Trans. Soc. Rheo!., 14 (1970) 555.
167. Leonov, A.I., Malkin, A.Ya. and Vinogradov, G.V., On the effect of rigidity
of dynamometers on the results of rheological measurements. Kolloid. J., 26
(1964) 335 (in Russian).
168. The Weissenberg Rheogoniometer. Instruction manual, Model R-18, Sangamo Controls Ltd, Bognor Regis, UK.
169. Leonov, A.I., A linear model of the stick-slip phenomena in polymer flows in
rheometer. Rheol. Acta, 23 (1984) 591.
170. Chernyak, Yu.B. and Leonov, A.I., On the theory of the adhesive friction of
elastomers. Wear, 108 (1986) 105.
171. Leonov, A.I., On the dependence of friction force on sliding velocity in the
theory of adhesive friction of elastomers. Wear, 141 (1990) 137.
172. Faitelson, L.A. and Alekseenko, A.I., Normal stresses at periodic shearing.
Mech. Polym., 2 (1973) 321 (in Russian).
173. Meissner, J., Modifications of the Weiss en berg rheogoniometer for measurement of transient rheological properties of molten polymers under shear.
Comparison with tensile data. J. Appl. Polym. Sci., 16 (1972) 2877.
174. Sakai, M., Fukaya, H. and Nagasama, M., Time dependent viscoelastic
properties of concentrated polymer solutions. Trans. Soc. Rheol., 16 (1972)
635.
175. Leonov, A.I., Lipkina, E.Kh., Paskhin, E.D. and Prokunin, A.N., On
theoretical and experimental studies of shear deformations in elastic polymer liquids. Izvestiya A.N. SSSR, Mekhanika Zidkosti i Gasa, 3 (1975) 3 (in
Russian).
176. Vinogradov, G.V. and Prozorovskaya, N.V., The studies of polymer melts
on capillary rheometer of constant pressure. Plasticheskie Massi, 5 (1964) 50
(in Russian).

References

443

177. Ivanova, L.I., The study of viscous properties of rubbers by means of


capillary rheometry. PhD Thesis, Institute of Petrochemical Synthesis,
USSR Academy of Science, Moscow, 1968.
178. Malkin, AYa. and Chalykh, A.Ye., Diffusion and Viscosity of Polymers,
Khimiya, Moscow, 1979 (in Russian).
179. Bagley, E.B., End correction in the capillary flow of polyethylene. J. Appl.
Phys., 28 (1957) 624.
180. Giesekus, R., Nicht-lineare effekte beim stromen viskoelastischer flussigkeiten durch schlitz- und lochdusen. Rheol. Acta, 7 (1968) 127.
181. Paskhin, E.D., Motion of polymer liquids under unstable conditions in
channel terminal. Rheo!. Acta, 17 (1978) 663.
182. Mooney, M., Explicit formulas for slip and fluidity. J. Rheol., 2 (1931) 210.
183. Vinogradov, G.V. and Ivanova, L.I., Wall slippage and elastic turbulence of
polymers in rubbery state. Rheol. Acta, 7 (1968) 243.
184. Kargin, V.A. and Sogolova, T.I., On the method of studying the true flow
process in polymers. J. Phys. Chern., 23 (1949) 540 (in Russian).
185. Kargin, V.A. and Sogolova, T.A., The study of viscous flow of polyisobutylene. J. Phys. Chern., 23 (1949) 550 (in Russian).
186. Vinogradov, G.V., Leonov, A.I. and Prokunin, AN., An uniaxial extension
of an elasto-viscous cylinder. Rheol. Acta, 8 (1969) 482.
187. Leonov, A.I., Prokunin, AN. and Vinogradov, G.V., Slow extension of
polymers as viscoelastic media. In Advances in Rheology of Polymers (Ed.
G.V. Vinogradov), Moscow, Khimiya, 1970, p. 41 (in Russian).
188. Ballman, R.L., Extensional flow of polystyrene melt. Rheol. Acta, 4 (1965)
137.
189. Meissner, J., Rheometer zur untersuchung der deformations mechanischen
eigenschaften von kunstoff-Schmelzen unter zugbeannspruchung. Rheol.
Acta, 8 (1969) 78.
190. Meissner, J., Dehnungsverhalten von Polyathylen-Schmelzen. Rheol. Acta,
10 (1971) 230.
191. Fikhman, Y.D., Radushkevich, B.V. and Vinogradov, G.V., Rheological
properties of polymers at extension with constant elongation rate and
elongation velocity. In Advances in Rheology of Polymers (Ed. G.V. Vinogradov), Moscow, Khimiya, 1970, p. 9 (in Russian).
192. Akutin, M.S., Prokunin, AN., Proskurnina, N.G. and Sabsay, O.Yu., On
extension of polymer melts with constant strain rate. Mech. Polym., N2
(1977) 353 (in Russian).
193. Prokunin, AN. and Proskurnina, N.G., On the role of rheology in the
process of extension of polymer melts under constant force. lnzenernoFizicheskii J., 36 (1979) 504 (in Russian).
194. Cogswell, F., Tensile deformations in molten polymers. Rheol. Acta, 8 (1969)
187.
195. Meissner, J., Development of a universal extensional rheometer for the
uniaxial extension of polymer melts. Trans. Soc. Rheol., 16 (1972) 405.
196. Donelly, G.J. and Weinberger, C.B., Stability of isothermal spinning of
Newtonian fluid. Ind. Chern. Fundamentals, 14 (1975) 334.
197. Prokunin, A.N. and Sevruk, V.D., On non-isothermal retardation of elastic
liquids. Inzenerno-Fizicheskii J., 43 (1982) 70 (in Russian).

444

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

198. Astarita, G. and Nicodemo, L., Extensional flow behavior of polymer


solutions. Chem. Eng. J., 1 (1970) 57.
199. Pokrovsky, V.N., Kruchinin, N.G., Danilin, G.A. and Serkov, A.T., A
relation between the coefficient of shear and longitudinal viscosities for
concentrated polymer solutions. Mech. Polym., 1 (1973) 124 (in Russian).
200. Kulichikhin, V.G. and Malkin, A.Ya., The rheology of uniaxial extension for
polymer systems. General principles and methods of study. Khimicheskie
Volokna, 1 (1975) 16 (in Russian).
201. Ziabicky, A., Fundamentals of Fibre Formation, Wiley, New York, 1976.
202. Han, CD., Rheology in Polymer Processing, Academic Press, New York,
1976.
203. Brewster, D., On some properties oflight. Phil. Trans. Roy. Soc., 103 (1813)
101.
204. Maxwell, lC, On the equilibrium of elastic solids. Trans. Roy. Soc.
Edinburgh, 20 (1850) 87.
205. Kuske, A., Photoelastic Stress Analysis, John Wiley, New York, 1974.
206. Kuhn, W. and Griin, F., Beziehungen zwischen elastischen Konstanten and
Dehnungsdoppelbrechnung hochelastischer Stoffe. Kolloid Z., 101 (1943)
248.
207. Upadhyay, R.K., Isayev, A.I. and Shen, S.F., Modeling of stresses in
multi-step shear deformation of polymeric fluids. J. Rheol, 27 (1983) 155.
208. Prokunin, A.N., Isayev, A.I. and Lipkina, E.Kh., Parallel superposition of
oscillatory motion on the steady flow of polymeric liquids. Mekhanika
Polimerov, 13 (1977) 589 (in Russian).
209. Osaki, K., Kimura, S. and Kurata, M., Relaxation of shear and normal
stresses in step-shear deformation of a polystyrene solution in comparison
with predictions of the Doi-Edwards. J. Polym. Soc., Polym. Phys. Ed., 19
(1981) 517.
210. Upadhyay, R.K., Isayev, A.1. and Shen, S.F., Transient shear flow behavior
of polymeric fluids according to the Leonov model. Rheol. Acta, 20 (1981)
443.
211. Vinogradov, G.V., Malkin, A.Ya., Plotnikova, E.P. and Kargin, V.A., On
the thixotropy of polymers in viscous state. Dokladi AN SSSR, 154 (1964)
1421 (in Russian).
212. Trapenznikov, A.A., On the limits of fluidity, critical elastic deformation and
critical strain rate of relaxing colloidal systems. Dokladi AN SSSR, 102
(1955) 1177 (in Russian).
213. Trapeznikov, A.A., On some ideas and methods of study of mechanical
properties for structural colloidal systems. Colloid J., 18 (1956) 438 (in
Russian).
214. Leonov, A.I., On the rheology of filled polymers. J. Rheol., 34 (1990)
1039.
215. Vinogradov, G.V. and Malkin, A.Ya., Rheological properties of polymers in
fluid state. Prikladnaya Mekhanika i Tekhnich. Fizika, 3 (1964) 66 (in
Russian).
216. Malkin, A.Ya., Yarlikov, B.Y. and Vinogradov, G.V., On the measurements
of relaxation properties of polymer melts at their deformation. Dokladi AN
SSSR, 174 (1967) 1370 (in Russian).

References

445

217. Vinogradov, G.V., Malkin, A.Ya. and Shumsky, V.F., Rheological properties of butyl rubber and low molecular polyisobutylene. In Advances in
Polymer Rheology (Ed. G.V. Vinogradov) Khimia, Moscow, 1970, p. 206 (in
Russian).
218. Lodge, A.S. and Meissner, S., Comparison of network theory prediction
with stress/time data in shear and elongation for a low-density polyethylene
melt. Rheo!. Acta, 12 (1973) 41.
219. Gortemaker, F.H., Hansen, M.G., de Cindio, B., Janeschitz-Kriegel, H., A
design of cone-and-plate apparatus for the measurement of the flow birefringence of polymer melts. Rheol. Acta, 15 (1976) 242.
220. Benbow, S.S. and Howells, E.R., Normal stress, shear recovery and viscosity
in polydimethylsiloxanes. Polymer, 2 (1961) 429.
221. Markovitz, H. and Williamson, R.B., Normal stress effect in polyisobutylene
solutions. Trans. Soc. Rheol., 1 (1957) 25.
222. Meissner, J., Die Kunstoffe-Schmelze als elastische Flussigkeit. 2.
Rheologische Verhalten in nichtlinearen Beanspruchungs beracht. Kunstoffe, 57 (1967) 702.
223. Ginn, R.F. and Metzner, A.B., Measurement of stresses developed in steady
laminar shearing flows of viscoelastic media. Trans. Soc. Rheol., 13 (1969) 429.
224. Sakamoto, K. and Porter, R.S., Comparison of normal stress measurements
on polymer melts. J. Polym. Sci. (Part 1B), 8 (1970) 177.
225. West, G.H., The application of molecular theories to the behavior of bulk
polymers in steady shear and recovery. Rheol. Acta, 9 (1970) 554.
226. Pritchard, W.G., Measurements of the viscometric functions for a fluid in
steady shear flow. Phil. Trans. Roy. Soc., 270 (1971) 507.
227. Higashitani, K. and Pritchard, W.G., A kinematic calculations of intrinsic
errors in pressure measurements made with holes. Trans. Soc. Rheol., 16
(1972) 687.
228. Tanner, R.I., Correlation of normal stress data for polyisobutylene solutions. Trans. Soc. Rheol., 17 (1973) 365.
229. Harris, E.K., Effect of blending on rheological properties of polystyrene. J.
Appl. Polym. Sci., 17 (1973) 1679.
230. Christiansen, E.B. and Leppard, W.R., Steady-state and oscillatory flow
properties of polymer solutions. Trans. Soc. Rheol., 18 (1974) 65.
231. Cox, W.P. and Merz, E.H.; Correlation of dynamic and steady flow
viscosities. J. Polym. Sci., 28 (1958) 619.
232. Lodge, A.S. and Meissner, J., On the use of instantaneous strains, superposed on shear and elongational flows of polymeric liquids, to test the
Gaussian network hypothesis and to estimate the segment concentratibn
and its variation during flow. Rheol. Acta., 11 (1972) 351.
233. Osaki, K., Tamura, M., Kurata, M. and Kotaka, T., Complex modulus of
concentrated polymer solutions in steady shear. J. Phys. Chem., 69 (1965)
4183.
234. Boij, H.C., Influence of superimposed steady shear flow on the dynamic
properties of non-Newtonian fluids. Rheo!. Acta,S (1966) 215.
235. Kotaka, T. and Osaki, K., Normal stresses, non-Newtonian flow and
dynamic mechanical behavior of polymer solutions. J. Polym. Sci., Part C,
15 (1966) 453.

446

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

236. Williams, M.e., Bird, RB., Oscillatory behaviour of normal stresses in


viscoelastic fluids. Eng. Chern. Fundamentals, 3 (1964) 42.
237. Simmons, 1.M., Dynamic modulus of polyisobutylene solutions in the
relationships among variations measures of fluid elasticity. J. Polym. Sci.,
A-2, 9 (1971) 1703.
238. Laufer, Z., Jalink, H.L. and Staverman, A.S., Dynamic properties of some
239.
240.
241.
242.
243.
244.

polymer solutions subjected to a steady shear superimposed on an oscillatory shear flow. I. Experimental results. Rheol. Acta, 14 (1975) 641.
Philippoff, W., Vibrational measurements with large amplitudes. Trans. Soc.
Rheol., 10 (1966) 317.
Briedis, I.P., Yakovlev, Yu.G. and Faitelson, L.A., Variations in viscoelastic
characteristics of highly filled plasticated rubber under periodic shearing.
Mekhanika Polimerov, 2 (1968) 29 (in Russian).
Faitelson, L.A. and Tsiprin, M.G., Dependence of viscoelastic properties of
polymer melts on filler loading, frequency and oscillation shear rate.
Mekhanika Polimerov, 3 (1968) 515 (in Russian).
Akers, L.e. and Williams, M.e., Oscillatory normal stresses in dilute
polymer solutions. J. Chern. Phys., 51 (1969) 3834.
Faitelson, L.A., Leonov, A.I. and Tsiprin, M.G., Periodic deformations and
viscous flow of polymer solutions and melts. Mekhanika Polimerov, 1(3)
(1969) 519; 2(2) (1970) 357; 3(3) (1970) 521 (in Russian).
Endo, H. and Nagasawa, M., Normal stress and shear stress in viscoelastic liquid under oscillatory shear flow. J. Polym. Sci., Part A-2, 8 (1970)

371.
245. Tanner, RI. and Simmons, 1.M., Combined simple and sinusoidal shearing
in elastic liquids. Chern. Eng. Sci., 22 (1967) 1803.
246. Isayev, A.I. and Hieber, e.A., Oscillatory shear flows of polymeric systems.
J. Polym. Sci., Polym. Phys. Ed., 20 (1982) 423.
247. Isayev, A.I. and Wong, e.M., Parallel superposition of small- and largeamplitude oscillations upon steady shear flow of polymer fluids. J. Polym.
Sci., Part B: Polym. Phys., 26 (1988) 2303.
248. Malkin, A.Ya., Isayev, A.I., and Vinogradov, G.V., Energetic evaluation of

efficiency of vibration actions in the flow of polymeric and filled systems.


Mekhanika Polimerov, 11 (1975) 439 (in Russian).
249. Pipkin, A. e., Small displacements superposed on visco metric flow. Trans.
Soc. Rheol, 12 (1968) 397.
250. Pipkin, A.e. and Owen, D.R., Nearly viscometric flows. Phys. Fluids, 10

(1967) 836.
251. Bernstein, B. and Huilgol, RR., On the ultrasonic dynamic viscosities in
superposed oscillatory shear. Trans. Soc. Rheol., 15 (1971) 731.
252. Bernstein, B. and Huilgol, RR, On ultrasonic dynamic moduli. Trans. Soc.
Rheol., 18 (1974) 583.
253. Powell, RL. and Schwarz, W.H., Rheological properties of aqueous poly
(ethylene oxide) solutions in parallel superposed flows. Trans. Soc. Rheo!., 17
(1975) 617.
254. Powell, RL. and Schwarz, W.H., Geometrical effects in the measurements of
mechanical properties in parallel superposed flows. J. Polym. Sci., Polym.
Phys. Ed., 17 (1979) 969.

References

447

255. Powell, R.L. and Schwarz, W.H., Nonlinear dynamic viscoelasticity. J.


Rheol., 23 (1979) 323.
256. Simmons, I.M., Dynamic modulus of polyisobutylene solutions in superposed steady shear flow. Rheol. Acta, 7 (1968) 184.
257. Kwon, Y. and Leonov, A.I., Remarks on orthogonal superposition of small
amplitude oscillations on steady shear flow. Rheol. Acta 32 (1993) 108.
258. Wong, CM. and Isayev, A.I., Orthogonal superposition of small and large
amplitude oscillations upon steady shear flow of polymer fluids. Rheol. Acta,
28 (1989) 176.
259. Meissner, 1., Neue Messmoglichkeiten mit einem zur untersuchung von
Kunstoff-Schmelzen geeigneten modifizierten Weissenberg-Rheogoniometer. Rheol. Acta, 14 (1975) 201.
260. Vinogradov, G.V., Malkin, A.Ya. and Berezhnaya, G.V., Rheological properties of concentrated polymer solutions. Visokomolekularnye Soyedineniya,
13A (1983) 2793 (in Russian).
261. Trapeznikov, AA and Assonova, T.V., The study of strength and high
elastic properties of solutions of rubbers and their vulcanizates at high shear
rates. Kolloid. J., 20 (1958) 376 (in Russian).
262. Vinogradov, G.V. and Malkin, A.Ya., Correlation of polymer properties at
continuous and periodic deformations. In Advances in Polymer Rheology
(Ed. G.V. Vinogradov), Khimiya, Moscow, 1970, p. 52.
263. Stratton, R.A. and Butcher, A.F., Experimental determination of the relationships among various measures of fluid elasticity. J. Polymer Sci., 9 A-2
(1971) 1703.
264. Tsiprin, M.G. and Faitelson, L.A., The peculiarities of flow of LDPE
melt filled with calcium metal silicate. Mekhanika Polimerov, 4 (1969)
693.
265. Pollett, W.F.O., Nature of strain recovery in polyvinyl chloride following
steady laminar shearing. Rheol. Acta, 1 (1958) 257.
266. Kwon, Y. and Leonov, A.I., On ID instabilities in simple shear and
extensional flows as predicted by some Maxwell-like constitutive equations.
J. Rheol. 36(8) (1992) 1515.
267. Schleiniger, G. and Weinacht, R.I., A remark on the Giesekus viscoelastic
fluid. J. Rheol., 35 (1991) 1157.
268. Trouton, F.T., On the coefficient of viscous traction and its relation to that
of viscosity. Proc. Roy. Soc., S.A., 77 (1906) 426.
269. Nitschman, H. and Schrade, 1., Uber das Fadenzeihvermogen bei Flussigkeiten als Viscosittanomalie. Helv. Chem. Acta, 31 (1948) 297.
270. Petrie, C.J.S., Elongation Flows, Pitman, San Francisco, 1979.
271. Prokunin, A.N. and Proskurnina, N.G., On the rheology in extension of
polymeric fluids. Inzhenerno-Fizicheskii J., 36 (1979) 42 (in Russian).
272. Fikhman, V.D., Radushkevich, B.v., Goldina, E.G. and Vinogradov, G.V.,
Viscous properties of polymers in elongation flow under constant velocity of
stretching and constant strain rate. Mekhanika Polimerov, 1 (1974) 137 (in
Russian).
273. Laun, H.M. and Miinstedt, H., Comparison of the elongation behavior of a
polyethylene melt at constant stress and constant strain rate. Rheol. Acta, 15
(1976) 517.

448

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

274. Laun, H.M. and Miinstedt, H., Elongational behavior of a low density
polyethylene melt I. Rheol. Acta, 17 (1978) 415.
275. Miinstedt, H. and Laun, H.M., Elongational behavior of a low density
polyethylene melt II. Rheol. Acta, 18 (1979) 492.
276. Prokunin, A.N. and Filippova, N.P., Slowing down of flow of a polymer
liquid at large elastic deformations. Inzhenerno-Fizicheskii J., 37 (1979) 724
(in Russian).
277. Radushkevich, B.V., Fikhman, V.D. and Vinogradov, G.V., Viscous and
relaxation properties of polymers in extension. In Advances in Polymer
Rheology (Ed. G.V. Vinogradov), Khimiya, Moscow, 1970 (in Russian).
278. Stevenson, J.E., Elongational flow of polymer melts. AIChE J., 18 (1972) 540.
279. Vinogradov, G.V., Dreval, V.E., Borisenkova, E.K., Kurbanaliev, M.K. and
Shalganova, V.G., Elongation deformations and rupture of linear flexiblechain polymers. Visokomolekularnye Soyedineniya, A23 (1981) 2627 (in
Russian).
280. Malkin, A.Ya., Volosevich, V.V. and Vinogradov, G.V., Evaluation of
ultimate conditions for uniaxial drawing of fluid polymers. International
Symposium on Synthetic Fibers: Physical and Chemical Processes in Fiber
Formation, Section I., Preprint of Institute of Synthetic Fibers, 1974, p. 196.
281. Vinogradov, G.V., Ultimate regimes of deformation of linear flexible chain
fluid polymers. Polymer, 18 (1977) 1275.
282. Prokunin, A.N. and Sevruk, V.D., Demonstration of flow slowing down
effect at different regimes of extension. Inzhenerno-Fizicheskii J., 41 (1981) 74
(in Russian).
283. Slonimsky, G.L. and Musaelyan, I.N., The study of fluidity of polyisobutylene. Visokomolekularnye Soyedineniya, 6 (1964) 1001 (in Russian).
284. Briedis, 1.1. and Leitland, V.V., Elongation and shear periodic deformations
of polymers in viscous-fluid state. Mekhanika Polimerov, 5 (1972) 880 (in
Russian).
285. Leonov, A.I. and Prokunin, A.N., On nonlinear effects in extensional flow of
polymeric liquids. Rheol. Acta, 22 (1983) 137.
286. Pro kunin, A.N., Some methods of experimental studies in the extension of
polymeric liquids. Intern. J. Polym. Mater., 8 (1980) 303.
287. Prokunin, A.N. and Sevruk, Y.D., On relaxation processes in polymeric
liquids. Inzhenerno-Fizicheskii J., 42 (1982) 987 (in Russian).
288. Baranboim, N.K., Mechanochemistry of High Molecular Combinations, Moscow, Chimiya, 1978 (in Russian).
289. Goldberg, V.M., Yarlykov, B.V., Paverman, N.G., Berezina, E.!., Akutin,
M.S. and Vinogradov, G.Y., Mechanodegradation of low density polyethylene in flow. Visokomolekularnye Soedineniya, 20A (1978) 2437 (in Russian).
290. Leonov, A.I., Lipkina, E.Kh. and Prokunin, A.N., Theoretical description of
fluidity loss for polymeric liquids under their intense deformations. Prikladnaya Mekhanika i Technich. Fizika, 4 (1976) 86 (in Russian).
291. Vinogradov, G.V., Insarova, N.I., Boiko, B.B. and Borisenkova, E.K.,
Critical regimes of shear in linear polymers. Polymer Eng. Sci., 12 (1972) 323.
292. Vinogradov, G.V., Malkin, A.Ya., Yanovski, Yu.G., Yarlykov, E.K. and
Berezhnaya, G.V., Viscoelastic properties and flow of narrow distribution
polybutadienes and polyisoprenes. J. Polym. Sci., A2 (1972) 1061.

References

449

293. Vinogradov, G.V., Malkin, A.Ya., Blinova, N.K., Sergeyenkov, S.I.,


Zabugina, M.O., Titkova, L.V., Yanovski, Yu.G. and Shalganova, V.G.,
Peculiarities of flow and viscoelatic properties of solutions of polymers with
a narrow molecular-weight distribution. European Polymer, 9 (1973) 1231.
294. Leonov, A.I. and Prokunin, A.N., On the spinnability for a viscoelastic
liquid. Izvestiya AN SSSR, Mekhanika Zhidkosti i Gaza, 5 (1973) 24.
295. Prokunin, A.N., Krasovitsky, B.A. and Sevruk, Y.D., On a possible mechanism of flow slow down for polymeric liquids. Inzhenerno-Fizicheskii .T., 40
(1981) 46 (in Russian).
296. Upadhyay, R.K. and Isayev, A.I., Elongational flow behavior of polymeric
fluids according to the Leonov model. Rheol. Acta, 22 (1983) 557.
297. Ishikuza, O. and Koyama, K., Elongational viscosity at a constant elongational strain rate of polypropylene melt. Polymer, 21 (1977) 164.
298. Matsumoto, T. and Boque, D.C., Non-isothermal rheological response
during elongational flow. Trans. Soc. Rheol., 21 (1977) 453.
299. Acierno, D., La Mantia, F.P., de Cindio, B. and Nikodemo, L., Transient
shear and elongational data for polyisobutylene melts. Trans. Soc. Rheol., 21
(1977) 261.
300. Prokunin, A.N., On a description of motions of elastoviscous polymeric
media. Inzhenerno-Fizicheskii .T., 54 (1988) 221 (in Russian).
301. Bell, A.D., Zadvornykh, V.N., Pro kunin, A.N. and Sysoyev, V.I., On the
manifestations of small weight - large relaxation time mechanism in
motions of polymers. Izvestiya AN SSSR, Mekhanika Zhidkosti i Gasa, 6
(1988) 3 (in Russian).
302. Leonov, A.I. and Srinivasan, A., On the modeling of fluidity loss phenomena
in Couette and Poiseuille flows of elastic liquids. Rheol. Acta, 30 (1991) 14.
303. Vasilchenko, S.V. and Potapov, A.G., The effect of elastic properties of
structurized systems on the ascent of a gas bubble in these. Colloid .T. USSR,
2 (1989) 353 (in Russian).
304. Ericksen, J.L., Overdetermination of the speed in rectilinear motion of
non-Newtonian fluids. Quart. Appl. Math., 14 (1956) 318.
305. Green, A.E. and Rivlin, R.S., Steady flows of non Newtonian fluids through
tubes. Quart. Appl. Math., 14 (1956) 299.
306. Criminale Jr., W.O., Ericksen, J.L. and Filby, G.L., Steady shear flows of
non-Newtonian fluids. Arch. Rat. Mech. Anal., 1 (1957/58) 410.
307. Yin, W.L. and Pipkin, A.C., Kinematics of viscometric flow. Arch. Rat.
Mech. Anal., 37 (1970) 111.
308. Langlois, W.E. and Rivlin, R.S., Slow steady-state flow of visco-elastic fluids
through non-circular tubes. Rend. Mat., 22 (1963) 169.
309. Wheeler, J.A. and Wissler, E.H., Steady flows of non-Newtonian fluids in a
square duct. Trans. Soc. Rheol., 10 (1966) 353.
310. Ramacharyulu, N.C.P., Slow steady flow of and idealized elastico-viscous
incompressible fluid of Oldroyd type through a straight tube with an
arbitrary cross-section. Z. Angew. Math. Mech., 47 (1967) 359.
311. Giesekus, H., Sekundarstromungen in viskoelastischen Fliissigkeiten by
stationarer und periodischer Bewegung. Rheol. Acta, 4 (1965) 85.
312. Semionov, V.V., Sekundarstromungen hochpolymerer schmelzen in Rohr
for elliptischen Querschnitt. Rheol. Acta, 6 (1967) 171.

450

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

313. Dodson, A.G., Townsend, P. and Walters, K., Non-Newtonian flow in pipes
of non-circular cross-section. Comput. Fluids., 2 (1974) 317.
314. Overdiep, W.S. and Van Krevelen, D.W., Studies of non-Newtonian Flow. I.
Criterion of flow instability. J. Appl. Polym. Sci., 9 (1965) 2779.
315. Etter, I. and Schowalter, W.R, Unsteady flow of an Oldroyd fluid in a
circular tube. Trans. Soc. Rheol., 9 (1965) 351.
316. Fielder, R. and Thomas, RH., The unsteady flow of a lamina in an
elastico-viscous liquid. Rheol. Acta, 6 (1967) 306.
317. Waters, N.D. and King, M.J., Unsteady flow of an elastico-viscous liquid.
Rheol. Acta, 9 (1970) 345.
318. Waters, N.D. and King, M.l., The unsteady flow of an elastico-viscous
liquid in straight pipes of circular cross-section. J. Phys. D 14 (1971)
204.
319. Chong, 1.S. and Vezzi, D.M., Unsteady flow of viscoelastic fluids. J. Appl.
Polym. Sci., 14 (1970) 17.
320. Chong, 1.S. and Frank, G.E., Time-dependent flow of quasi-linear viscoelastic fluids. J. Appl. Polym. Sci., 14 (1970) 1639.
321. Akay, G., Numerical solutions of some unsteady laminar flows of viscoelastic fluids in concentric annuli with axially moving boundaries. Rheol. Acta.,
16 (1977) 598.
322. Duffy, B.R., Flow of a liquid with an anisotropic viscosity tensor. J.
Non-Newtonian Fluid Mech., 4 (1978) 177; 7 (1980) 107.
323. Ryan, M.E. and Dutta, A., Analysis of constant rate starting flow of a
viscoelastic fluid in annual, cylindrical and planar conduits. J. Rheol., 25
(1981) 193.
324. Isayev, A.I., Unsteady channel flow of polymeric fluids. J. Rheol., 28 (1984)
411.
325. Coleman, B.D. and Markovitz, H., Normal stress effects in second-order
fluids. J. Appl. Phys., 35 (1964) 1.
326. Isayev, A.I., Upadhyay, RK. and Shen, S.F., Experimental and theoretical
investigation of polymer flow in a converging and a diverging channel. SPE
ANTEC Tech. Pap., 28 (1982) 298.
327. Barhes, H.A., Townsend, P. and Walters, K., Flow of non-Newtonian
liquids under a varying pressure gradient. Nature, 224 (1969) 585.
328. Barhes, H.A., Townsend, P. and Walters, K., On pulsate flow of nonNewtonian liquids. Rheol. Acta, 10 (1971) 517.
329. Bassov, N.I., Leonov, A.I., Lubartovich, S.A. and Phelipchuk, 1.1., Nutzung
von vibrationsimpulsen bei der verarbeitung von polymeren. Plaste und
Kautschuk, 6 (1974) 436.
330. Bassov, N.I., Leonov, A.I., Lubartovich, S.A. and Phelipchuk, I.I., Unter-

suchung des vibrationspritzgiessprocess von hochpolymeren. I.


Rheologische eigenschaften von polymerschmelzen bei periodischen scher
und volumdeformation mit grossen amplituden. Plaste und Kautschuk, H12

(1975) 970.

331. Bassov, N.I., Leonov, A.I., Lubartovich, S.A. and Phelipchuk, 1.1., Untersuchung des vibrationspritzgiessprocess von hochpolymeren 2. Analyse des
processes der erzeugung periodischer druckimpulse. Plaste und Kautschuk,
H3 (1976) 198

References

451

332. Bassov, N.I., Leonov, A.I., Lubartovich, S.A. and Phelipchuk, 1.1., Unter-

suchung des vibrationspritzgiessprocess von hochpolymeren 3. Zur ausbreitung der longitudinal wellen in schmelzestrom. Plaste und Kautschuk, H7

(1977) 490.
333. Bassov, N.J., Lubartovich, S.A. and Lubartovich, V.A., Vibroforming of
Polymers, Khimiya, Leningrad, 1979 (in Russian).
334. Pozdeyev, V.A. and Shakirov, N.V., On pulsatile flow of viscoelastic liquids.
In News in Polymer Rheology (Ed. G.V. Vinogradov), Institute of Petrochemical Synthesis AN USSR, Moscow, 1985, pp. 232, 233.
335. Wein, O. and Sobolik, V., Periodic longitudinal flows of pseudoplastic
materials. 'Collection of Czechoslovak Chemical Communications, 45 (1980)
1010.
336. Ballenger, T.F. and White, J.L., An experimental study of flow patterns in
polymer fluids in the reservoir of a capillary rheometer. Chem. Eng. Sci., 25
(1970) 1191.
337. Bagley, E.B. and Birks, A.M., Flow of polyethylene into a capillary. J. Appl.
Phys., 31 (1960) 556.
338. Cogswell, F.N., Converging flow of polymer melts in extrusion dies. Polym.
Eng. Sci., 12 (1972) 64.
339. Clegg, P.L., The flow properties of polyethylene and their effect on fabrication.
Trans. J. Plast. [nst., 26 (1958) 151.
340. Den Otter, J.L., Mechanisms of melt fracture. Plastics Polym., 38 (1970) 1555.
341. Oyanagi, Y. and Yamaguchi, Y., J. Soc. Mater. Sci. Jap. 20 (1971) 659.
342. Oyanagi, Y., Appl. Polym. Symp., 20 (1973) 123.
343. Schott, H. and Kaghan, W.S., Flow irregularities in the extrusion of
polyethylene melts. Ind. Eng. Chem., 51 (1959) 844.
344. Tordella, J.P., Capillary flow of molten polyethylene - a photographic study
of melt fracture. Trans. Soc. Rheol., 1 (1957) 203.
345. Tordella, J.P., Unstable flows of molten polymers: a second site of melt
fracture. J. Appl. Polym. Sci., 7 (1963) 215.
346. Brizitsky, V.I., Vinogradov, G.V., Isayev, A.I. and Podolsky, Y.Y. Extensional
stresses during polymer flow in ducts. J. Appl. Polym. Sci., 22 (1978) 751.
347. Ballenger, T.F., Chen, J.J., Crowder, J.W., Hagler, G.E., Bogue, D.e. and

White, J.L., Polymer melt flow instabilities in extrusion: investigation of the


mechanism and material and geometrical variables. Trans. Soc. Rheol., 15

(1971) 195.
348. White, J.L. and Condo, A., Flow patterns in polyethylene and polystyrene

melts during extrusion through a die entry region: measurements and


interpretation. J. Non-Newt. Fluid. Mech., 3 (1978-79) 41.
349. Ma, e.Y., White, J.L., Weissert, F.e. and Min, K., Flow patterns in carbon
black filled polyethylene at the entrance to a die. J. Non-Newt. Fluid Mech.,

17 (1985) 275.
350. White, S.A. and Baird, D.G., The importance of extensional flow properties
on planar entry flow patterns of polymer melts. J. Non-Newt. Fluid Mech.,
20 (1986) 93.
351. Han, e.D. and Drexler, L.H., Studies of converging flows of viscoelastic

polymeric melts. I. Stress-Birefringent measurements in the entrance region


of a sharp-edged slit die. J. Appl. Polym. Sci., 17 (1973) 2329.

452

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

352. Evans, RE. and Walters, K., Flow characteristics associated with abrupt
changes in geometry in the case of highly elastic liquids. J. Non-Newt. Fluid.
Mech., 20 (1986) 11.
353. Boger, D.V., Viscoelastic flows through contractions. Ann. Rev. Fluid.
Mech., 19 (1987) 157.
354. Miinstedt, H., Dependence of the elongational behavior of polystyrene melts
on molecular weight and molecular weight distribution. J. Rheo/., 24 (1980)
847.
355. Upadhyay, R.K. and Isayev, A.I., Simulation of two-dimensional planar flow
of viscoelastic fluid. Rheol. Acta., 25 (1986) 80.
356. Lawler, J.V., Muller, S.G., Brown, R.A. and Armstrong, RC., Laser Doppler
velocimetry measurements of velocity fields and transitions in the viscoelastic liquids. J. Non-Newt. Fluid Mech., 20 (1986) 51.
357. Ames, W.F., Numerical Methodsfor Partial Differential Equations, 2nd edn.,
Nelson, London, 1977.
358. Crochet, M.J., Davis, A.R and Walters, K., Numerical Simulations of
Non-Newtonian Flows, Elsevier, Amsterdam, 1984.
359. Keuning, R., Simulation of viscoelastic fluid flow. In Fundamentals of
Computer Modeling for Polymer Processing, (Ed. c.L. Tucker III), Hanser,
Munich, 1989, p. 404.
360. Keuning, R., On the high Weissenberg number problem. J. Non-Newt.
Fluid. Mech., 20 (1986) 209.
361. Beris, A.N., Armstrong, RC. and Brown, RA., Finite element calculation of
viscoelastic flow in a journal bearing: I. Small eccentricities. J. Non-Newt.
Fluid Mech., 16 (1984) 141.
362. Yeh, P.W., Kim-E., M.E., Armstrong, RC. and Brown, R.A., Multiple
solutions in the calculation of axisymmetric contraction flow of an Upper
Convected Maxwell fluid. J. Non-Newt. Fluid Mech., 16 (1984) 173.
363. Shulman, Z.P. and Khusid, B.M., Non-steady Processes of Convective
Transfer in Hereditary Media, Minsk, Nauka i Tekhnika, 1984 (in
Russian).
364. Astarita, G. and Sarti, G.S., The dissipative mechanism in flowing polymers:
theory and experiments. J. Non-Newt Fluid Mech., 1 (1976) 39.
365. Astarita, G. and Sarti, G.S., Thermodynamics of compressible materials
with entropic elasticity. In Theoretical Rheology (Ed. J.F. Hutton, J.RA.
Pearson and K. Walters), Wiley, New York, 1974, p. 123.
366. Van Krevelen, D.V., Properties of Polymers. Correlation with Chemical
Structure, Elsevier, New York, 1972.
367. Pakharenko, V.A., Zverev, V.G., Privalko, V.P., Petrushenko, E.F.,
Demchenko, S.S. and Gladchenko, A.N., Thermophysical and Rheological
Characteristics and Friction Coefficients of Filled Thermoplastics, (Ed. Yu.S.
Lipatov), Naukova Dumka, Kiev, 1983 (in Russian).
368. Tomer, RV., Theoretical Foundations of Polymer Processing, Khimiya
Moscow, 1977 (in Russian).
369. Encyclopedia of Polymers, Vol. 2, Soviet Encyclopedia, Moscow, 1974, p.
1032 (in Russian).
370. Vargaftic, N.B., Reference Book of Thermophysical Properties of Gases and
Liquids, Nauka, Moscow, 1972 (in Russian).

References

453

371. Sedov, L.I., Introduction in Continuum Mechanics, Fizmatgiz, Moscow, 1962


(in Russian).
372. Braun, H. and Friederich, Chr., Transient processes in Couette flow of a
Leonov Liquid influenced by dissipation. J. Non-Newt. Fluid Mech., 33
(1989) 39.
373. Alexandrov, A.P. and Lazurkin, Yu.S., Study of polymers: high elastic
deformations of polymers. J. Techn. Phys., 9 (14) (1939) 1249 (in Russian).
374. Markovitz, H., Superposition in rheology. J. Polym. Sci., 50 (1975) 431.
375. Vinogradov, G.V., Malkin, A.Ya., Yanovsky, Yu.G., Dzyura, E.A., and
Shumsky, V.F., Rheological properties of polymers in fluid state. Mekhanika
Polimerov, 1 (1969) 164 (in Russian).
376. Stolin, AM., Bostandjian, S.A. and Plothikova, N.V., Critical conditions for
hydrodynamic thermal explosion in flow of power liquid. Heat and Mass
Transfer in Rheologically Complex Systems, 7 (1976) 261 (in Russian).
377. Prokunin, A.N., Zadvornykh, V.N. and Sysoyev, V.I., On shear self-heated
flow of viscoelastic liquid between two rotating discs. Inzhenerno-Fizicheskii
J., 54 (1988) 282 (in Russian).
378. Frank-Kamenetsky, D.A, Diffusion and Heat Transfer in Chemical Kinetics,
Nauka, Moscow, 1967 (in Russian).
379. Merzhanov, A.G. and Stolin, AM., To the thermal theory of flow of viscous
liquid. Dokladi AN SSSR, 198 (1971) 1294 (in Russian).
380. Bostandjian, S.A., Merzhanov, A.G. and Khudyaev, S.I., On the hydrodynamic thermal explosion. Dokladi AN SSSR, 163 (1965) 133 (in Russian).
381. Bostandjian, S.A. and Stolin, A.M., On the critical conditions of thermal
regime of a generalized Couette flow. Inzhenerno-Fizicheskii J., 17 (1969) 86
(in Russian).
382. Kaganov, S.A., On steady laminar flow of an incompressible liquid in flat
and circular tube with account of dissipative heat and dependence of
viscosity on temperature. Prikladnaya Mekhanika i Tekhnich. Fizika, 3
(1962) 96 (in Russian).
383. Merzhanov, A.G., Posecelskii, A.P., Stolin, A.M. and Shteinberg, A.S.,
Experimental accomplishment of the hydrodynamic thermal explosion.
Dokladi AN SSSR, 210 (1973) 52 (in Russian).
384. Prokunin, A.N., Polymeric liquid sealing of a rotating shaft and rheological
effects accompanying its functioning. Rheol. Acta., 29 (1990) 231.
385. Bukhman, Yu.A., Lipatov, V.N., Prokunin, AN., Soboleva, L.I., Sysoyev,
V.I. and Shumilov, v.P., A new type of sealing device based on the normal
stress effect. In Improvements in Equipment and Technology of Drilling for Oil
and Gas, AlI-Union Research Institute of Drilling Technology (VNIIBT),
Issue 57 (1983) 42 (in Russian).
386. Prokunin, A.N., Sysoyev, V.I., Bukhman, Yu.A. and Zadvornykh, V.N., The
sealing of rotating shaft with a polymeric liquid. In Pumping Aggregates and
Fittings for Hydrotransport of Solid Materials, AlI-Union Institute Hydro
Pipe-Line (GIDROTRUBOPROVOD), 1987, p. 92 (in Russian).
387. Bukhman, Yu.A., Dagayev, G.F., Entov, V.M., Konstantinov, I.L., Leonov,
A.I., Lipatov, V.N., Prokunin, A.N. and Shumilov, v.P., A sealing of
rotating shaft. USSR Certificate of Invention (Internal USSR Patent) No
731 156 1980 (in Russian).

454

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

388. An Tuan, Nguen and Sadradze, G.V., A study of hydrostatic thrust bearing
with a viscoelastic lubricant. Bull-AN Georgia SSR, 72 (1973) 129 (in
Russian).
389. Galakhov, M.A., Physico-mathematical foundations of elasto-hydrodynamic
theory of lubricating. Preprint of Inst. Probl. Mech. (IPM), Moscow, AN
SSSR, No. 94 (1977) (in Russian).
390. Merzhanov, A.G. and Stolin, A.M., Hydrodynamic analogies of the ignition
and extinction phenomena. Prikladnaya Mekhanika i Tekhnich. Fizika, 1
(1974) 65 (in Russian).
391. Vulis, L.A. and Dzhaugashtin, K.E., Magneto-gas-dynamic Couette flow. J.
Techn. Phys., 34 (1964) 2171 (in Russian).
392. Shukhov, V.G., Pipe lines and their application to the oil industry. Russkoe
Tovarischestvo Pechati, Moscow, 1895 (in Russian).
393. Targ, C.M., Basic Problems of the Theory of Laminar Flow, Gostekhizdat,
Moscow, 1951 (in Russian).
394. Stolin, A.M. and Merzhanov, A.G., Quasi-steady thermal regime in a head
flow of liquid. In Transaction of IV All-Union Meeting on Heat- and Mass
Transfer. Inst. Heat and Mass Transfer, AN Belorussk. SSR, 3 (1972) 341 (in
Russian).
395. Vulis, L.A., Thermal Regime of Combustion, Gosenergoizdat, Moscow, 1954
(in Russian).
396. Upadhyay, R.K. and Isayev, A.I., Nonisothermal elongational flow of
polymeric fluids according to the Leonov model. J. Rheol., 28 (1984) 581.
397. Matsui, M. and Bogue, D.C., Studies in non-isothermal rheology. Trans.
Soc. Rheol., 21 (1977) 133.
398. Carey, D.A., Wust, Jr., c.J. and Bogue, D.C., Studies in non-isothermal
rheology: behavior near the glass transition temperature and in the oriented
glassy state. J. Appl. Polym. Sci., 25 (1980) 575.
399. Racin, R. and Bogue, D.C., Molecular effects in die swell and shear
rheology. J. Rheol., 23 (1979) 263.
400. Prokunin, A.N. and Sevruk, V.D., On the swelling effect at withdrawal of an
elastic fluid out of capillary. Inzhenerno-Fizichesky J., 34 (1980) 343 (in
Russian).
401. Rivlin, R.S., Solution of some problems in the exact theory of viscoelasticity.
J. Rational Mech. Anal., 5 (1956) 179.
402. Vinogradov, G.V., Mamakov, A.A. and Pavlov, V.N., Experimental study of
anamalous-viscous bodies under complex stressed state. Izvestiya AN SSSR,
Mekhanika i Mashinostroyenie 6 (1959) 100 (in Russian).
403. Vinogradov, G.V., Mamakov, A.A. and Tyabin, N.V., The flow of anomalous-viscous bodies under complex stressed state. Izvestiya AN SSSR,
Mekhanika i Mashinostroyenie, 2 (1960) 65 (in Russian).
404. Pasley, P.R. and Slibar, A., Laminar flow of drilling mud due to axial
pressure gradient and external torque. Petroleum Trans. AIME, 210 (1957)
310.

405. Coleman, B.D. and Noll, W., Helical flow of general fluids. J. Appl. Phys.,
30 (1959) 1508.
406. Fredrickson, A.G., Helical flow of an annular mass of viscoelastic fluid.
Chem. Eng. Sci., 11 (1960) 252.

References

455

407. Tanner, R.I., Non-Newtonian flow and oil seal problem. J. Mech. Eng. Sci.,
2 (1960) 21.
408. Gutkin, A.M., Punching a viscoplastic disperse mass between sliding parallel plates. Colloid. J., 23 (1961) 20 (in Russian).
409. Myasnikov, V.P., Complex shear flow of visco-plastic liquid. Prikladnaya
Mekhanika i Tekhnicheskaya Fizika, 5 (1961) 98 (in Russian).
410. Dierckes, A.C. and Schowalter, W.R., Helical flow of non-Newtonian
polyisobutylene solution. Ind. Eng. Chern. Fund., 5 (1966) 263.
411. Bortnikov, V.G., Kuznetsov, E.V. and Tyabin, N.V., Extrusion of polyethylene tubes with using rotating forming elements. Plasticheskiye Massi, 8
(1967) 49 (in Russian).
412. Tyabin, N.V., Bortnikov, V.G., Tsentovsky, E.M. and Vachagin, K.D., Flow
of polymer melts in rotating extruder head. Mekhanika Polimerov, 3 (1968)
531 (in Russian).
413. Prokunin, A.N., Fridman, M.L. and Vinogradov, G.V., Helical flow of
non-Newtonian polymer media. Mekhanika Polimerov, 3 (1971) 497 (in
Russian).
414. Bukhman, Yu.A., Zadvornykh, V.N., Litvinov, A.I. and Shulman, Z.P.,
Helical flow of nonlinear viscoplastic media in annular die. InzenernoFizicheskii J., 49 (1985) 221 (in Russian).
415. Kazakia, J.Y. and Rivlin, R.S., The influence of vibration on Poiseuille flow
of non-Newtonian fluid. Rheol. Acta, 17 (1978) 210.
416. Manero, 0., Meka, B. and Valenzuela, R., Further developments of nonNewtonian flow in oscillating pipes. Rheol. Acta, 17 (1978) 693.
417. Lyubartovich, S.A., Lyubartovich, V.A. and Stolbov, V.V., A theoretical
analysis of polymer extrusion through a vibrational head. Theor. Found.
Chern. Technol., 18 (1984) 561 (in Russian).
418. Prokunin, A.N., To the theory of disc extrusion. Izvestiya AN SSR,
Mekhanika Zhidkostu i Gaza, 1 (1972) 169 (in Russian).
419. Maxwell, B. and Scalor, A.I., The elastic melt extruder - works without
screw. Mod. Plastics, 37 (1959) 107.
420. Anikin, A.S., Kim, V.S., Korneyev, Yu.A. and Makarov, V.L., A laboratory
disc extruder. In Processing of Polymer Materials (Ed. K.A. Salazkin), Vol.
35, MIKhM, Moscow 1970 p. 10 (in Russian).
421. Ostapchuk, Yu.G., Disc Extruders, Tekhnika, Kiev, 1972 (in Russian).
422. Smelkov, P.E. and Kozulin, N.A., Elasticity of polymer melts and its
practical application. J. Appl. Chern., 35 (1962) 2693 (in Russian).
423. Yartsev, N.K., Knyazkin, I.P. and Sporyagin, E.A., The work of disc
extruder with screw feeder. Plasticheskie Massi, 9 (1968) 44 (in Russian).
424. Leonov, A.I., On optimal geometry for generalized rotational flow of
elastoviscous liquids in thin gaps. Ann. N. Y. Acad. Sci., 491 (1987) 130.
425. Huilgol, R.R., On generalized torsional flow. Trans. Soc. Rheol., 18 (1974) 191.
426. Yanovski, Yu.G. and Malkin, A.Ya., The correlation of polymer properties
at continuous and periodic deformations. In Advances in Polymer Rheology
(ed. G.V. Vinogradov) Khimiya, Moscow, 1970, p. 52 (in Russian).
427. Prokunin, A.N. and Proskurnina, N.G., On the relationship between uniform and non-uniform extensions of viscoelastic liquids. InzhenernoFizicheskii J., 34 (1978) 629 (in Russian).

456

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

428. Litvinov, V.G., Extension of viscoelastic jet. Mekhanika Polymerov, 2 (1967)


326 (in Russian).
429. Prokunin, A.N. and Proskurnina, N.G., On the mechanics of steady
withdrawal of polymeric extrudate from capillary. Khimicheskie Volokna, 5
(1978) 22 (in Russian).
430. Winter, H.H. A collaborative study on the relation between film blowing
performance and rheological properties of two low-density and two highdensity polyethylene samples. Pure App/. Chem., 55 (1983) 943.
431. Sevruk, V.D., Prokunin, A.N. and Fridman, M.L., Regularities in Extension
of Polyer Melts and Their Manifestations in Plastics' Processing, VIINTEKHIM, Moscow, 1984 (in Russian).
432. Fridman, M.L., Sevruk, V.D., Prokunin, A.N. and Novikov, D.D., Estimation of quality of thermoplastics for agriculture films. Plasticheskie Massi, 7
(1984) 61 (in Russian).
433. Leonov, A.1. and Prokunin, A.N., On the stretching and swelling of an
elastic liquid extruded from a capillary die. Rheo/. Acta, 23 (1984) 62.
434. Whipple, B.A. and Hill, c.T., Velocity distribution in die swell. AIChE J., 24
(1978) 664.
435. Chen, I., Hagler, G.E., Abbott, L.E., Bogue, D.C. and White, J.L., Interpretation of tensile and melt spinning experiments on low density and high
density polyethylenes. Trans. Soc. Rheol., 16 (1972) 473.
436. Prokunin, A.N. and Proskurnina, N.G., Theoretical and experimental study of
extension of viscoelastic liquids. In Machines and Technology of Processing of
Rubbers, Rubber Mixtures and Polymers, Yaroslavil', 1977, p. 52 (in Russian).
437. Malkin, A.Ya., Goncharenko, V.V. Malinovskii, V.V., Swelling of polymer
jet extruded from capillary and slit dies. Mekhanika Polymerov, 3 (1976) 487
(in Russian).
438. Grassley, W.W., Glasscock, S.D. and Crawley, RL., Die swell in molten
polymers. Trans. Soc. Rheo/., 14 (1970) 519.
439. Denn, M.M., Petrie, c.J.S. and Avenas, P., Mechanics of steady spinning of
viscoelastic liquids. AIChE J., 21 (1975) 791.
440. James, P.F., Open channel siphon with viscoelastic liquids. Nature, 212
(1977) 754.
441. Prokunin, A.N., A model of elastic deformation for the description of
withdrawal of polymer solutions. Rheol. Acta, 22 (1983) 374.
442. Leonov, A.I., Propagation of nonlinear waves along a viscoelastic bar.
Rheol. Acta, 27 (1988) 335,
443. Antman, S.S., Nonuniquiness of equilibrium states for bars in tension. J.
Math. Anal. Appl., 44 (1973) 333.
444. Antman, S.S., Qualitative theory of ordinary differential equations of nonlinear elasticity. In Mechanics Today, Vol. 1. (Ed. S. Nemat-Nasser), Pergamon, New York, 1974, p. 58.
445. Bland, D.R, Nonlinear Dynamic Elasticity, Blaisdall, London, 1969.
446. Coleman, B.D., Gurtin, M.E., Herrera, I. and Truesdell, R and c., Wave
Propagation in Dissipative Materials, Springer, New York, 1965.
447. Barenblatt, G.I., On propagation of instantaneous disturbances in a medium with a nonlinear stress-strain relation. PMM: J. Appl. Math. Mech., 17
(1953) 14 (in Russian).

References

457

448. Kestenbaum, H.S., Roslyakov, G.S. and Chudov, L.A., Point Explosion:
Methods of Calculation: Tables, Nauka, Moscow, 1974 (in Russian).
449. Meissner, J., Stephenson, S.E., Dermamels, A. and Portmann, P., Multiaxial
elongational flows of polymer melts: classification and experimental realization. J. Non-Newtonian Fluid Mech., 7 (1982) 10.
450. Garvey, B.S., Whitlock, M.H. and Freese, J.A., Processing characteristics of
synthetic tire rubber, Ind. Eng. Chern. 34 (1942) 1309.
451. Nason, H.K., A high temperature, high pressure rheometer for plastics. J.
Appl. Phys., 16 (1945) 338.
452. Spencer, R.S. and Dillon, R.E., The viscous flow of molten polystyrene - I.
J. Coil. Sci., 3 (1948) 163.
453. Spencer, R.S. and Dillon, R.E., The viscous flow of molten polystyrene - II.
J. Coli. Sci., 4 (1949) 241.
454. Spencer, R.S., Flow oflinear amorphous polymers. J. Polyrn. Sci., 5 (1950) 591.
455. Tordella, J.P., Fracture in the extrusion of amorphous polymer through
capillaries. J. Appl. Phys., 27 (1956) 454.
456. Tordella, J.P., An instability in the flow of molten polymers. Rheo/. Acta, 1
(1958) 216.
457. Tordella, J.P., Unstable flow of molten polymers: a second site of melt
fracture. J. Appl. Polymer Sci., 7 (1963) 215.
458. Tordella, J.P., An unusual mechanism of extrusion of polytetrafluoroethylene at high temperature and pressure. Trans. Soc. Rheol., 7 (1963) 231; US
Patent 2 791 806 (1967).
459. Tordella, J.P., Unstable flow of molteD polymers. In Rheology, Vol. V (ed.
F.R. Eirich), Academic Press, New York, 1969.
460. Clegg, P.L., The flow of molten polymer and their effect on fabrication. Brit.
P/ast., 30 (1957) 535.
461. White, J.L., Critique of flow pattern in polymer fluids at the entrance of a
die and instabilities leading to extrudate distortion. App/. Polym. Symp., 20
(1973) 155.
462. Petrie, c.J.S. and Denn, M.M., Instabilities in polymer processing. AIChE
J., 22 (1976) 209.
463. Boudreaux, Jr., E. and Cuculo, J.A., Polymer flow instability: a review and
analysis. J. Macro. Sci., CI6(1) (1977) 39.
464. Bagley, E.B., The separation of elastic and viscous effects in polymer flow.
Trans. Soc. Rheol., 5 (1961) 355.
465. Ballenger, T.F. and White, J.L., The development of the velocity field in
polymer melts in a reservoir approaching a capillary die. J. App/. Polym.
Sci., 15 (1971) 1949.
466. Bialas, G.A. and White, J.L., Extrusion of polymer melts and melt flow
instabilities. I. Experimental study of capillary flow and extrudate distortion.
Rubber Chem. Techno!., 42 (1969) 675.
467. Bialas, G.A. and White, J.L., Extrusion of polymer melts and melt flow
instabilities. II. Site of initiation and mechanisms of melt flow instability.
Rubber Chem. Technol., 42 (1969) 682.
468. Schreiber, H.P., Bagley, E.B. and Birks, A.M., Filament distortion and die
entry angle effects in polyethylene extrusion. J. Appl. Polym. Sci., 4 (1960)
362.

458

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

469. Dennison, M.T., Flow instability in polymer melts: a review. Plastics


Polymers, 35 (1967) 803.
470. Ramamurthy, A.V., Wall slip in viscous fluids and influence of materials of
construction. Trans. Soc. Rheol., 18 (1972) 431.
471. Yung, K.M., Melt instability in super-rate extrusion. MS Thesis, North
Carolina State University, Raleigh, 1973.
472. El Kissi, N. and Piau, J.M., J. Non-Newt. Fluid Mech., 37 (1990) 55.
473. Ramamurthy, A.V., LLDPE rheology and blown film fabrication. Advances
in Polymer Technology, 6(4) (1986) 489.
474. Lupton, J.M. and Regester, R.W., Melt flow of polyethylene at high rates.
Polym. Eng. Sci., 5 (1965) 235.
475. Metzner, A.B., Carley, E.L. and Park, I.K., Polymer melts: a study of
steady-state flow, extrudate irregularities and normal stresses. Modern
Plastics, 133 (July 1960) 133.
476. Kurtz, S.J., Die geometry solutions to sharkskin melt fracture. In Advances
in Rheology (Proc. IX Int. Congress Rheology), Vol. 3. (Ed. B. Mena, A.
Garcia-Rejon and C. Rangel-Nafaile), University of National Autonoma
Mexico, Mexico City, 1984, p. 399.
477. Brydson, J.A., Elastic effects in polymer melt flow, In Flow Properties of
Polymer Melts, Van Nostrand Reinhold Co., New York, 1970, p. 79.
478. Cogswell, F.N., Polymer Melt Rheology, Halsted Press, Wiley, New York, 1981.
479. Herranen, M. and Savolaninen, A., Correlation between melt fracture and
ultrasonic velocity. Rheol. Acta, 23 (1984) 461.
480. Bagley, E.B., Cabot, I.M. and West D.C., Discontinuity in the flow curve of
polyethylene. J. Appl. Phys., 29 (1958) 109.
481. Uhland, E., Das anomale flieBverhalten von polyathylen hoher dichte,
Rheol. Acta, 18 (1979) 1.
482. Myerholtz, R.W., Oscillating flow behavior of high-density polyethylene
melts. Appl. Polym. Sci., 11 (1967) 687.
483. Metzger, A.P., Hamilton, C.W. and Merz, E.H., Anomalous flow behavior of
high density polyethylene melts. S.P.E. Trans., 3(1) (1963) 21.
484. Kalika, D.S. and Denn, M.M., Wall slip and extrudate distortion in linear
low-density polyethylene. J. Rheol., 31 (1987) 815.
485. Lim, F.J. and Schowalter, W.K., Wall slip of narrow molecular weight
distribution polybutadienes. J. Rheol., 33 (1989) 1359.
486. Bagley, E.B. and Schreiber, H.P., Paper presented at the 31st Annual
Meeting of the Society of Rheology, Pittsburgh, PA, November 1960.
487. Bagley, E.B. and Schreiber, H.P., Effect of die entry geometry on polymer
melt fracture and extrudate distortion. Trans. Soc. Rheol., 5 (1961) 341.
488. Cogswell, F.N., The influence of pressure on the viscosity of polymer melts.
Plastics Polymers, 41 (1973) 39.
489. White, J.L. and Kondo, A., Flow patterns in polyethylene and polystyrene
melts during extrusion through a die entry region: measurement and
interpretation. J. Non-Newt. Fluid. Mech., 3 (1977) 41.
490. Benbow, J.J., Brown, R.N. and Howells, E.R., The flow of elastic liquids at
room temperature. Phenomenes de relaxation et de fluage en rheologie
Non-linearire, Centre National de la Recherche Scientifique, Colloque
XCVII, Paris, France, 1961, p. 65.

References

459

491. Benbow, 1.1., Charley, R.V. and Lamb, P., Unstable flow of molten polymers. Nature, 192 (1961) 223.
492. Howells, E.B. and Benbow, 1.1., Flow defects in polymer melts. Plastics
Polymers, 30 (1962) 240.
493. Clegg, P.L., Flow in various extrusion processes. Brit. Plast., 39 (1966) 96.
494. Cogswell, F.N., Polymer melt rheology during elongational flow. Appl.
Polym. Symp., 27 (1975) 1.
495. Cogswell, F.N., Stretching flow instabilities at the exits of extrusion dies. J.
Non-Newt. Fluid Mech., 2(1) (1977) 37.
496. Clegg, P.L., Die design in extrusion. Trans. Plast. Inst., 8 (1960) 245.
497. Clegg, P.L. and Huck, N.D., Effect of extrusion variables on the fundamental properties of tubular polyethylene film. S.P.E. Trans., 1 (1961)
121.
498. Stabler, H.G., Haward, R.V. and Wright, B., The flow properties of
high-density polyethylene in capillary rheometry and extrusion. In Advances
in Polymer Science and Technology, Society of the Chemical Industry,
London, 1967, p. 327.
499. Pearson, 1.R.A., Mechanisms of melt flow instability. Plastics Polymers, 37
(1969) 285.
500. Vinogradov, G.V., Flow and rubber elasticity of polymeric systems. Pure
and Appl. Chem., 26 (1971) 423.
501. Vinogradov, G.V., Critical regimes of deformation of liquid polymeric
systems. Rheol. Acta, 12 (1973) 273.
502. Vinogradov, G.V., Viscoelastic and fracture phenomena in uniaxial extension of high-molecular linear polymers. Rheol. Acta, 14 (1975) 942.
503. Vinogradov, G.Y., Limiting regimes of deformation of polymers. Polym.
Eng. Sci., 21 (1981) 339.
504. Vinogradov, G.V., Isayev, A.I. and Katsyutsevich, E.V., Critical regimes of
505.
506.
507.
508.

509.

510.
511.

oscillatory deformation of polymeric systems above glass transition and


melting temperatures. J. Appl. Polym. Sci., 22 (1978) 727.
Vinogradov, G.V. and Manin, Y.N., An experimental study of elastic
turbulence. Koll Z., 201 (1965) 93.
Vinogradov, G.V., Friedman, M.L., Yarlykov, N.V. and Malkin, A.Ya.,
Unsteady flow of polymer melts: polypropylene. Rheol. Acta, 9 (1970) 323.
Vinogradov, G.V., Viscoelasticity of polymeric systems in fluid and rubbery
states in uniaxial extension and shear. Pure and Appl. Chem., 42 (1975) 527.
Vinogradov, G.V., Protasov, P. and Dreval, Y.E., The rheological behavior
of flexible-chain polymers in the region of high shear rates and stresses, the
critical process of spurting, and supercritical conditions of their movement
at T> I'g. Rheol. Acta, 23 (1984) 46.
Vinogradov, G.V., Yanovsky, Yu.G., Titkova, L.V., Barancheeva, V.V.,
Sergeenkov, S.I. and Borisenkova, E.K., Viscoelastic properties of linear
polymers in the fluid state and their transition to the high-elastic state.
Polym. Eng. Sci., 20(17) (1980) 1138.
Hatzikiriakos, S.G. and Dealy, 1.M., Wall slip of molten high density
polyethylene. I. Sliding plate rheometer studies. J. Rheol., 35(4) (1991) 497.
Hatzikiriakos, S.G. and Dealy, 1.M., The effect of interface conditions on
wall slip and melt fracture of high density polyethylene. In Proc. 49th

460

512.
513.
514.
515.
516.
517.
518.
519.
520.
521.
522.
523.
524.
525.
526.
528.

529.
530.
531.
532.

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

Technical Coriference of the SPE (ANTEC), Society of Plastic Engineers,


Brookfield, CT, 1991, p. 2311.
Ramamurthy, A.V., US Patent 4552 712 (1985).
Ramamurthy, A.V., US Patent 4554 120 (1985).
Ramamurthy, A.V., US Patent 4522 776 (1985).
Piau, J.M., El Kissi, N. and Tremblay, B., Low Reynolds number flow
visualization of linear and branched silicones upstream of orifice dies. J.
Non-Newt. Fluid Mech., 30 (1988) 197.
Piau, J.M., El Kissi, N. and Tremblay, B., Influence of upstream instabilities
and wall slip on melt fracture and sharkskin phenomena during silicones
extrusion through orifice dies. J. Non-Newt. Fluid Mech., 34 (1990) 145.
Metzner, A.B., White, J.L. and Denn, M.M., Behavior of viscoelastic
materials in short-time processes. Chem. Eng. Prog., 62 (1966) 81.
Metzner, A.B., White, J.L. and Denn, M.M., Constitutive equations for
viscoelastic fluids for short deformation periods and for rapidly changing
flows: significance of the Deborah number. AIChE J., 12 (1966) 863.
Malkin, A.Ya. and Leonov, A.I., Unstable flows of polymers. In Advances in
Rheology of Polymers (Ed. G.V. Vinogradov), Moscow, Khimiya, 1970, p. 98
(in Russian).
Denn, M.M., Issues in viscoelastic fluid mechanics. Ann. Rev. Fluid. Mech.,
22 (1990) 13.
Hutton, J.F., Fracture of liquids in shear. Nature, 200 (1963) 646.
Hutton, J.F., The fracture of liquids in shear: the effects of size and shape.
Proc. R. Soc. London, Series A, 287 (1965) 222.
Bagley, E.B., Rudin, A. and Schreiber, H.P., Fracture of liquid in shear.
Nature, 203 (4941) (1964) 175.
Hills, D.A., Hasegava, T. and Denn, M.M., On the apparent relation
between adhesive failure and melt fracture. J. Rheol., 34 (1990) 891.
Benbow, J.J. and Lamb, P., New aspects of melt fracture. S.P.E. Trans., 3
(1963) 7.
Clegg, P.L., Elastic effects in the extension of polyethylene. In The Rheology
of Elastomers (Ed. P. Mason and N. Wookey), Pergamon Press, New York,
1958, p. 174.
Borisenkova, E.K, Dreval, V.E., Vinogradov, G.V., Kurbanaliev, M.K,
Moiseyev, V.V. and Shalganova, V.G., Transition of polymers from the fluid
to the forced high elastic and leathery states at temperatures above the glass
transition temperature. Polymer, 23 (1982) 91.
Bulgin, D., Hubbard, G.D. and Walters, M.H. Road and laboratory studies
of friction of elastomers In Proc. 4th Rubber Technol. Conf., Institute of the
Rubber Industry, London, 1962, p. 173.
Grosh, KA., The relation between the friction and viscoelastic properties of
rubbers. Proc. Roy. Soc. London, A274 (1963) 21.
Bartenev, G.M. and Yelkin, A.I., On the mechanism of friction of high
elastic materials at high and low temperatures. In The Theory of Friction
and Wear, Nauka, Moscow, 1965, p. 95 (in Russian).
Frenkin, E.1. and Yanovsky, Yu.G., Viscoelasticity of polymers and their
behavior in friction. In Advances in Rheology of Polymers (Ed. G.V.
Vinogradov), Khimiya, Moscow, 1970, p. 269 (in Russian).

References

461

533. Schallamach, A., A theory of dynamic rubber friction. Wear, 6 (1963) 375.
534. Schallamach, A., A theory of dynamic rubber friction. Rubber Chem.
Technol., 39 (1966) 320.
535. Barguins, M., Sliding friction of rubber and Schallamach waves- a review.
Mat. Sci. Eng., 73 (1985) 45.
536. Barguins, M. and Roberts, A.D., Rubber friction variation with rate and
temperature: some new observations. J. Phys. D: Appl. Phys., 19 (1986)
547.
537. Roberts, A.D. and Thomas, A.G., Static friction of smooth clean vulcanized
rubber. Rubber. Chem. Techn. 50 (1977) 266.
538. Leonov, A.I. and Srinivasan, A., On self-oscillating of an elastic plate sliding
over a smooth surface. Int. J. Eng. Sci., 31 (1993) 453.
539. Karakin, A.V. and Leonov, A.I., On self-oscillations at capillary polymer
flows. Pridkladnaya Mekhanika i Technicheskaya Fizika, 3 (1968) 110 (in
Russian).
540. Buyevich, Yu.A. and Leonov, A.I., Self-oscillations in Couette flow of
incompressible Maxwellian liquid. Prikladnaya Mekhanika i Technicheskaya
Fizika, 2 (1966) 103 (in Russian).
541. Pearson, lR.A. and Petrie, C.J.S., On the melt flow Instability of Extruded
Polymers. In Polymer Systems, Deformation and Flow (Ed. R.E. Wetton and
R.W. Whorl ow), Macmillan, London, 1968, p. 163.
542. Renardy, M., Short wave instabilities resulting from memory slip. J.
Non-Newt. Fluid Mech., 35 (1990) 73.
543. Adewale, K.E.P. and Leonov, A.I., On modeling spurt flows of polymers. J.
Non-Newt. Fluid Mech., 49 (1993) 133.
544. Smith, T., Strength and extensibility of elastomers. In Rheology (Ed. F.R.
Eirich), Vol. 5, Academic Press, New York, 1969, p. 160.
545. Metzner, A.B., Extensional primary field approximation for viscoelastic
media. Rheol. Acta, 10 (1971) 434.
546. Sridhar, T. and Gupta, R.K., Fluid detachment and slip in extensional flows.
J. Non-Newt. Fluid Mech., 30 (1988) 285.
547. Pearson, lR.A. and Matovich, M.A., Spinning molten polymers. Ind. Eng.
Chem. Fund., 8 (1969) 605.
548. Miller, J.C., Swelling behavior at extrusion. SPE Trans., 3 (1963) 134.
549. Ishihava, H. and Kase, S., Studies of melt spinning. V. Draw resonance as a
limit cycle. J. Appl. Polym. Sci., 1 (1975) 557.
550. Ishihava, H. and Kase, S., Studies of melt spinning. VI. Simulations of draw
resonance using Newtonian and Power Law viscosities. J. Appl. Polym. Sci.,
20 (1976) 169.
551. Weinberger, e.B., Cruz-Saenz, G.F. and Donnelly, G.J., Onset of draw
resonance during isothermal melt spinning: a comparison between measurements and predictions. AIChE J., 22 (1976) 441.
552. Gelder, D., The stability of fiber drawing processes. Ind. Eng. Chem.
Fundam., 10 (1971) 534.
553. Hyun, J.e., Theory of draw resonance. Part I: Newtonian Fluids; Part II:
Power Law and Maxwell Fluids. AIChE J., 24 (1982) 418.
554. Berman, V.S. and Yarin, A.L., Dynamic regimes of filament formation.
Izvestiya AN SSSR, Mekhanika Zhidkosti i Gaza 6 (1983) 31 (in Russian).

462

Nonlinear Viscoelastic Effects in Flows of Polymer Melts

555. Varin, A.L., Onset of self-oscillations at filament formation. PMM: Appl.


Math. Mech., 47 (1983) 82 (in Russian).
556. Levich, V.G., Physico-Chemical Hydrodynamics, Fizmatgiz, Moscow, 1959

(English Translation).

557. Goldin, M., Yerushalmi, J., Pfeffer, R. and Shinnar, R., Breakup of a laminar
capillary jet of a viscoelastic fluid. J. Fluid Mech., 38 (1969) 689.
558. Rubin, H. and Wharshavsky, M., A note on the break-up of viscoelastic
liquid jet. Isr. J. Technol., 8 (1970) 285.
559. Middleman, S., Stability of viscoelastic jet. Chem. Eng. Sci., 20 (1965) 1037.
560. Bazilevsky, A.V., Entov, Y.M. and Rozhkov, A.N., Elastic stresses in
capillary jets of dilute polymer solutions. Izvestiya AN SSSR, Mekhanika
Zhidkosti i Gaza, 2 (1985) 3 (in Russian).
561. Entov, V.M. and Yarin, A.L., On the effect of elastic stresses on capillary
breakup of jets of dilute polymer solutions. Izvestiya AN SSSR, Mekhanika
Zhidkosti i Gaza, 1 (1984) 27 (in Russian).
562. Hinch, E.J., Mechanical models of dilute polymer solutions in strong flows.
Phys. Fluids, 20 (1977) 30.
563. Boussfield, D.W., Kennings, R., Marucci, G. and Denn, M.M., Nonlinear
564.
565.
566.
567.
568.
569.
570.
571.

572.

analysis of the surface tension driven breakup of viscoelastic filaments. J.


Non-Newt. Fluid Mech., 21 (1986) 79.
Lenezyk, S.P. and Kiser, K.M., Stability of vertical jets of non-Newtonian
fluids. AIChE J., 17 (1971) 826.
Isui, H. and Sano, Y., Elongational flow of dilute drag reducing fluids in a
falling jet. Phys. Fluids, 24 (1981) 214.
Slonimsky, G.L., Askadsky, A.A. and Logvinenko, V.K., On some peculiarities in flow of polyisobutylene. Mekhanika Polymerov, 4 (1967) 959 (in
Russian).
Eder, G., Janeschitz-Kriegl, H., and Liedauer, S., Crystallization processes in
quiescent and moving polymer melts under heat transfer conditions. Progr.
Polym. Sci., 15 (1990) 629.
Van der Vegt, A.K. and Smith, P.P.A., Crystallization phenomena in flowing
polymers. Soc. Chem. Ind. London Monographs, 26 (1967) 313.
Mackley, M.R. and Keller, A., Flow induced crystallization of polyethylene
melt. Polymer, 14 (1973) 16.
Vinogradov, G.V., Malkin, A.Ya., Borisenkova, E.K., Yarlykov, B.V. and
Kargin, V.A., On a possible mechanism of melt flow instability. Dokladi AN
SSSR, 179 (1968) 1366 (in Russian).
Grassie, N. and Scott, N., Polymer Degradation and Stabilization, Elsevier,
New York, 1990.
Emmanuel, N.M. and Buchachenko, N.M., Chemical Physics of Polymer
Degradation and Stabilization, VNU Science Press, Utrecht, Netherlands,

1987.
573. White, J.L., Rheological behavior and boundary condition characteristics of
rubber compounds. J. Appl. Polym. Sci.: Appl. Polym. Symp., 50 (1992) 109.
574. Han, Ch.D., Multiphase Flow in Polymer Processing, Academic Press, New
York, 1981.
575. Doi, M. and Edwards, S.F., The Theory of Polymer Dynamics, Oxford Press,
New York, 1986.

References

463

576. Leonov, A.I., Basov, N.I. and Kazankov Yu.V., The Foundations of
Thermoreactive Resin and Rubber Processing by the Injection Molding
Method, Khimiya, Moscow, 1977 (in Russian).
577. Biesenberger, J.A. and Sebastian, D.H., Principles of Polymerization Engineering, Wiley, New York, 1983.
578. Schouten, J.A., Tensor Analysis for Physicists, Clarendon Press, Oxford,
1951.
579. Sokolnikoff, I.S., Tensor Analysis, Theory and Applications to Geometry and
Mechanics of Continua, 2nd edn, Wiley, New York, 1964.
580. Lurie, A.I., Nonlinear Theory of Elasticity, Nauka, Moscow, 1980 (in
Russian).
581. Grassley, W.W., The entaglement concept in polymer rheology. Adv. Polym.
Sci., 16 (1974) 1.
582. Borisenkova, E.K., Sabsai, O. Yu., Kurbanaliev, M.K., Dreval V.E., Vinogvadov, G.V. Criterion of ultimate strength of fluid high polymers in uniaxial
extension, Polymer 19 (1978) 1473.

Author Index

Bagley, E.B. 96, 357, 360, 364, 368,


443, 451, 457, 458, 460
Ballenger, T.F. 359, 368, 451
Barguins, M. 383, 461
Bartenev, G.M. 30, 437, 460
Benbow, 1.1. 370, 458-60
Beris, A.N. 37, 438
Bialas, G.A. 359, 457
Bird, R.B. 46, 53, 434, 440
Brewster, D. 110, 444
Brydson, 1.A. 361, 458
Cerf, R. 46, 439
Chandrasekhar, S. 46, 439
Clegg, P.L. 357, 382, 457, 460
Cogswell, F.N. 247, 361, 368, 370,
451, 458, 459
Curtiss, S.F. 53, 434, 440, 459
Dealy, 1.M. 373, 375, 388, 458
de Gennes, P.G. 48, 51, 439, 440
Denn, M.M. 368, 375, 376, 458, 460
Dennison, M.T. 359, 458
den Otter, 1.L. 368, 451
Dillon, R.E. 356, 357, 381, 457
Doi, M. 51, 52, 440

Dupret, F. 60, 63, 76, 441


Eckart, C. 6, 9, 436
Edwards, B.l. 37, 438
Edwards, S.F. 51, 52, 440
Ericksen, 1.L. 33, 224, 449
Garvey, B.C. 356, 457
Giesekus, H. vi, 434, 443
Goddard, 1.D. 38, 438
Godunov, S.K. vi, 58, 60, 64, 434, 441
Gordon, R.l. 50, 437
Gotlib, Yu.Ya. 45, 439
Green, A.E. 33, 441, 449
Green, M.S. 33, 49, 440
Grmela, M. 37, 438
Griin, F. 112, 144
Hatzikiriakos, S.G. 373, 375, 388,459
Herranen, M. 361, 362, 458
Howells, E.B. 370, 458, 459
Hutton, 1.F. 381, 460
laneschitz-Kriegl, H. 112, 405, 434,
462
laumann, G. 40, 439

Author Index

Joseph, D.D. 60, 76, 439, 441


Kalika, D.S. 368, 375, 376
Kargin, V.A. 45, 404, 439, 443
Kirkwood, J.G. 46, 439
Kissi, El N. 364, 368, 377, 458, 460
Kondo, K 6, 436
Kuhn, W. 112, 444
Kurtz, S.J. 360, 361, 371, 372, 458
Kuvshinsky, Ye.V. 4, 435
Landel R.F. 17, 437
Larson, R.G. 33, 36, 52, 434, 438
Leonov, A.1. 434-7, 441-4, 446-51,
453, 455, 456, 460, 461
Lim, F.l 368, 458
Lodge, AS. 33, 49, 112, 437, 440
Lupton, J.M. 360, 364, 366, 372, 378,
382, 388, 458
Malkus, D.S. 76, 441
Marchal, lM. 60, 63, 76, 441
Maxwell, J.e. 110, 444
Metzger, AP. 365, 458
Metzner, AB. 360, 438, 458, 460, 461
Miller, e. 38, 438
Murnaghan F.D. 10, 436
Myerholtz, R.W. 364, 368, 370, 372,
389, 458
Nason, H.K. 356, 457
Neumann, 110
Ogden, R.W. 17,437
Oldroyd, J.G. 5, 33, 37, 436, 437
Overdiep, W.S. 382, 450
Paskhin, Ye.D. 249, 381, 382, 443
Pearson, J.R.A. 371, 390, 434, 459,
461
Phan-Thien, N. 50, 438
Piau, J.M. 364, 368, 377, 378,458,460
Prokunin, A.N. 436, 437, 442-4,
447-9, 453-6
Ramamurthy, A.V. 359, 375, 376,458,
460

465

Regester, R.W. 360, 364, 366, 372,


378, 382, 388, 458
Riseman, J. 46, 439
Rivlin, R.S. 33, 37,437,438,449, 454,
455
Roberts, AD. 383, 461
Rouse Jr., P.E. 46, 439
Rutkevich, I.M. 60, 441
Savolaninen, A. 361, 362, 458
Sawyers, KN. 37, 438
Schallamach, A 383, 384, 461
Schowalter, W.R. 50, 368, 437, 450,
458
Schreiber, H.P. 359, 368, 457, 458
Semdyanov, A.I. 307
Slonimsky, G.L. 45, 439, 462
Sogolova, T.I. 404, 443
Spencer, R.S. 356, 357, 381, 457
Tanner, R.I. 50, 434, 438, 446, 455
Tobolsky, AV. 33, 49, 440
Tordella, J.P. 357, 359, 364, 365, 382,
389, 451, 457
Tremblay, B. 377, 460
Truesdell, e. 65, 436, 438
Uhland, E. 364, 458
Valanis, Ke. 17,437
Van Krevelen, D.W. 382,450
Vinogradov, G.V. 372, 374, 375, 378,
382, 394, 395, 396, 399, 400, 434,
435,440-9,451,453,454,459,460,
462
Volkenstein, M.V. 45, 439
Weissenberg, K 4, 435
White, J.L. 359, 368, 434, 438, 451,
456-8, 460, 462
Yamamoto, M. 50, 440
Yung, KM. 360, 458
Zaremba, S. 40, 438
Zimm, B. 46, 439

Subject Index

Activation energy 20, 262, 265


Adhesive friction of elastomers 384
Adiabatic sound velocity 346
see also Nonlinear wave
propagation along a
viscoelastic bar
Adiabatic temperature 287
Affinity assumption 44
Arrhenius temperature dependence
20
'Bamboo effect', see Extrudate
appearance
Bagley plot, see Capillary
rheometer, entry and exit
correction
Barrus-effect, see Die swell
Bar-wall adhesion 349, 350
Bar-wall separation condition 349,
350
'Bar approximation' 343
see also Nonlinear wave
propagation along a
viscoelastic bar
Basic invariants of second-order
tensors 11, 55

Basis vectors 412


Bead-and-spring model 46, 48
Biopolymer (Kelsan), 0.2% aqueous
solution 224
Biot number 273, 283
'Boger fluid' - 0.1 % solution of
polyacrylamide in maltose
syrup/water solvent 246-48
Brownian force 46, 48, 49
Buckling of polymer extrudate 381
Butyl rubber 267, 316
Butyl rubber, 40% solution in
transformer oil 119, 120, 125,
128, 135, 136, 138, 151, 157,
158, 160, 249, 251, 252
Capillary decay (break-up) 403
Capillary rheometer
scheme 94
flow curve 95
entry and exit correction 95-6
dissipative heating 96
wall slip 96-7
Cauchy deformation tensor 40
Cauchy strain tensor 8
total 8, 39

Subject Index

reversible 8
Causality principle 40
see also Rational mechanics
Cayley-Hamilton identity 42, 417
Cayley polar decomposition
theorem 339, 414
Certification of raw material 325
Chemorheology 409
Clausius-Duhem inequality 4, 58
Closure approximations 49
Cohesion rupture of polymer melts
382
Compatibility conditions 416
Component-wise expressions
momentum balance 426-33
continuity equation 426-8
temperature variations 428-9
velocity gradient tensor 429-33
upper convected time derivative
430-33
Compressibility relations for elastic
solids
general 15
linear 15
Cone-plate rheometer
basic relations 81--4
schemes and methods 84-7
methodical remarks 87-92
Configuration space 47
convective diffusion 47
Configuration tensor 56
as internal state parameter 35, 55
molecular sense 47, 48
evolution equation 58
positive definiteness 68-9
Conformation entropy 261
'Conjuncture' regions 246
see also Entrance steady flows
Conservation laws 423
Constitutive equations 1, 426
for uniform extension 188-9
for simple shear 114-7
Constitutive equations of integral
type
Rivlin-Sawyers 37
Goddard-Miller 38
Oldroyd-Lodge 38, 73
Kaye-BKZ class 38, 73--4

467

Wagner type 38, 73


Continuity equation 8, 422
Continuum law of motion 411
Contraction entrance flow 246
Cooling rate 291
Copolymers
ethylene-propylene 359
tetrafluoroethylene-hexafluoropropylene 359, 364
'Cork flow', see Extrudate
appearance
Couette plain flow 80, 272, 274
see also Simple shear
Couette rheometer, see Cylindercylinder rheometer
Creep function 28
Critical shear rate 361
Critical wall shear stress 360, 363,
364, 373
Cross-sectional heat transfer 283
Crosslinked rubbers 1, 2
Curie principle 426
Cylinder-cylinder rheometer 81
Deborah number 2, 117,381
Deformation history 39, 42
Density 418
Detachment of polymers from the
walls 281, 400--402
under extrudate stretching
400-401
under low-molecular liquid action
401-2
Diadics 413
Die swell, see Extrudate swell
Discontinuous flow curve 366
see also Spurt melt fracture
Disipative losses in swelling 337
see also Extrudate swell
Disk-disk rheometer 92
torque and thrust 92-3
scheme 92
Disk extrusion 305
slow flow 306-8
strong flow 308-9
Dissipative instability 54, 58-60
Dissipative function 13-14, 425
Dissipative heat generation 260-61,

468

Subject Index

Dissipative heat generation-contd.


264, 275, 282
Dissipative heat source 283
Dissipative media 423
Dissipative relation 55
Distribution function
equilibrium 43
non-equilibrium 44, 47
Gaussian 44
Inverse Langevin 45
Divinyl-nitrile rubber 249, 250
Doppler shift 62, 63
Drawing a polymer solution from
reservoir 338-43
Drawing a polymer from capillary
324
Drawing resonance 401
Droplet-thread structure 402-4
see also Capillary decay
Dumbbell model 46
Dynamic deformations
measurements 85
frequency 85, 118, 135-7
storage modulus 85, 118, 135-7
loss modulus 85, 118, 135-7
dynamic viscosity 85
phase shift 135-7
Eckart's concept of 'variable elastic
state' 5-6
see also 'Unloading in small'
principle
Effect of fluoroelastomer on stickslip 374-5
see also Spurt melt fracture
Effective extensional viscosity 177,
200-201
Effective shear viscosity 127
'Elastic limit' 10
Elastic potential 11
stability constraints 66
Elastic potential, specifications
Mooney-Rivlin 17, 66
of network theory 17
Valanis and Landell 17
Ogden power-like 17
BST 18
Power-like BST 18

Elastic shear modulus 2, 4


Elastic strain tensor, see
Recoverable strains
'Elastic turbulence', see Melt flow
instabilities
Elementary wave parameters
phase 62
frequency 62
wave vector (numbers) 62
End-to-end distance of
macromolecule 43
Entanglements 2, 11
Entrance steady flows
experiments 246-50
vortices 256
elongation effects 246, 247
numerical simulations 250-57
Entropy 11
single chain 44
Boltzman's 44
Gibbs' 45
Entropy balance equation 425
Entropy elasticity assumption
260-64
Entropy flux 425
Entropy production 14, 262, 425
Entry region of melt fracture
initiation 368
see also Melt flow instabilities
Equation for temperature variations
261
Eulerian coordinates 6, 39
Eulerian description of continuum
410
Euring's molecular mechanism 21,
30
Evolution equation for total strains
421
Exit flow 250, 252
Exit region of melt fracture
initiation 369
see also Melt flow instabilities
Expansion entrance flow 256
Extension shock waves 347
see also Nonlinear wave
propagation along a
viscoelastic bar
Extinction angle 132-3

Subject Index

Extrudate appearance 357, 358, 365


Extrudate stretching and swelling
326-37
Extrudate swell 296, 327, 357
Fading memory 27, 41
see also Rational mechanics
Failure envelope 399
Failure of adhesion 376
FENE approximation 50
FENE dumbbell model 150, 71
Finger strain tensor 6, 415
total 6, 8
reversible 6, 8, 14, 16
Finite element method (FEM) 252
First Law of Thermodynamics 262,
423
First normal stress difference 79,
116, 277
Floating thrust bearing 318-19
Flow-induced anisotropy 15, 50
Flow-induced crystallization
in simple extension 404
in capillary entrance 405
in simple shear 405
Flow-rate controlled capillary
rheometer 364
Flow-stress history 77
Flow between rotating disks
data 158
theory 159-60
Flow birefringence 109-112, 132,
240-41, 257-9
Flow curve 127
Flow rate 228, 237
Flows in channels and pipes 224
Fluid mechanical instability 380-81
see also Melt flow instabilities,
mechanisms
Fluidity loss 212-13, 382
Fluorocarbons
DFL spray 374
DYNAMAR 9613 374
Fokker-Plank equation 47
see also Statistical mechanics
Fourier thermal conductivity 262
Frechet theorem 41
see also Rational Mechanics

469

Fracture of polymer melts 381


see also Melt flow instabilities,
mechanisms
Frame of reference
fixed 410
movable 411
Free energy 10, 55, 261, 262
Free swelling, see Extrudate swell
Frozen coefficients 60-61
see also Hadamard instability
Galilean invariance 425
Gaussian chain 44, 46
Generalized rotational flow 310-13
median surface 310
curvature radii 310
surface of revolution 310
'gently sloping' 311
thrust functional 312
torque functional 312
gap volume functional 312
variational problems 313
torque extremum 313-15
thrust maximization 315
plate biconical geometry 317
Generalized thermodynamic fluxes
426
Generalized thermodynamic forces
425
Gibbs' relation 14, 424
Giesekus model 35, 72, 162-5,
197-8
Glycerin/water solution, 70% 403
Gordon-Schowalter/J ohnsonSegalman model 34, 63
Green (Almansi) strain tensor 415
Green measure of strain 415
Gross irregularities, see Extrudate
appearance
Hadamard instability
general 12, 54, 347
physical sense 76-7
differential models 60, 63-4
single integral models 67-8
Hardening phenomena
experimental 20, 30, 199, 201-10,
324-5

470

Subject Index

Hardening phenomena-contd.
modelling as relaxation transition
211-4, 218-21, 234--6
modelling with long time
relaxation mode 214-8, 221-3
Helical flow 298
Hencky strain tensor 10, 416
High-temperature flow regime
286-7, 288
High Deborah flows 253, 382
High density polyethylene, HDPE
247, 360, 363-5, 369, 373, 379,
407
High elastic modulus 27, 80, 265
Hookean spring force 44, 46
Hugoniot's adiabate 347
Hydrodynamic (drag) force
Stokes approximation 46
Oseen approximation 48
Hydrodynamic thermal explosion
274, 282
Hydrodynamic thermal ignition and
extinction 282
Hysteresis in flow curve 364
see also Spurt Melt fracture
Ill-posedness, see Hadamard
instability
Incompressibility condition 9, 418,
422
'Index of melt fluidity' 325
Inertialess approximation 237, 269
Instantaneous elastic modulus 27
Internal energy 11, 261, 423
Internal energy balance 423
Internal rotations 423
Irreversible thermodynamics, see
Non-Equilibrium
thermod ynamics
Irreversible strains 1
Isothermal sound velocity 346
see also Nonlinear wave
propagation along a
viscoelastic bar
Isotropic pressure 11
equilibrium 74--5
non-equilibrium 74-5
Isotropic tensor function 56, 416

Isotropy 11
Jaumann tensorial time derivative,
see Tensorial time derivative,
co-rotational
Kinematics of uniform uniaxial
extension
velocities and strain rate 97-8
stretching ratio 98
recoverable strain 99
irreversible longitudinal strain
rate 99
Kinetic coefficients 426
see also Non-equilibrium
thermodynamics
Kinetic energy 13, 423
Kinetic tensors
mobility 15, 56
thermal conductivity 15
see also Non-equilibrium
thermodynamics
Kuhn's segment 44
Lagrangean coordinates 6
Lagrangean description of
continuum 410
Land region of melt fracture
initiation 369
see also Melt flow instabilities
Langevin equation 46, 48, 49
see also Statistical mechanics
Long time relaxation mode 30, 123
see also Hardening phenomena,
modeling with long time
relaxation mode
Larson model 35, 72, 165-6, 197
Laser-Doppler anemometry 249-51
measurements 381, 395
Leonov model 35, 164
Linear low density polyethylene,
LLDPE 361-3, 371
Linear model of polymer wall slip
394
see also Melt flow instabilities,
mechanisms
Linear uniform extension
general relations 171

Subject Index

recoverable strain 172


start-up under given free end
velocity 172--4
Linear viscoelasticity 16, 25-8
Lip vortices 246
see also Entrance flows,
experiments
Local equilibrium assumption
423--4
Locality principle 40
see also Rational mechanics
Long-term durability of uncured
polymers 399-400
Loss of concavity in velocity profile
382
Low-temperature flow regime 285,
288
Low density polyethylene, LDPE,
119, 120, 127, 153, 101-3, 205,
207-20, 216, 247, 263, 267, 324,
358, 362, 368, 369, 372, 379,
407
Mach number 349
Macroscopic stress 44
Mass balance equation, see
Continuity equation
Mass conservation equation, see
Continuity equation
Master curve 266
Material objectivity 40
'Matte flow', see Extrudate
appearance
Maxwell-like constitutive equations
54,68
Maxwell liquids 55
Maxwell model
single modal 3, 4
multi-modal 4
Mean relaxation time 27, 265
Mechanical dissipation 264
Mechanical energy
balance 12, 13
flux 13
Melt flow instabilities 356
Melt fracture 357, 361, 362
mechanisms 379
molecular weight, MW, 379

471

molecular weight distribution,


MWD, 264, 369-70
polymer structure effect 369-70
Memory functional 27
see also Rational mechanics
Metric tensor 412-3
Mobility tensor, see Kinetic tensors,
mobility
Molecular weight between
entanglements 378
Momentum balance 423
Na-CMC (carboxymethyl cellulose)
0.25% aqueous solution 403
Neo-Hookean relation 62
Network
temporary 2, 49
Gaussian 49
New viscoelastic model 167-9, 198
Newtonian liquid 260--61, 273, 274
Newtonian rheology 273, 283
Newtonian viscosity 27, 80, 265,
278
Nitril rubber 383
Non-equilibrium thermodynamics 5,
14, 54, 56, 426
Non-evolutionarity, see Hadamard
instability
Non-isothermal flows 260--61
capillary flows 282
elongation flows 290--94
high De flows 282
Non-isothermal strain recovery
(retardation) 294-6
Nonlinear uniform extension
under constant strain rate 176-9,
189-90
under constant stress 179, 192-3
under constant force 179-81,
190--92
recovery strain 181-5, 193-5
stress and strain relaxation 185-8
stress - recoverable strain
envelope 183
Nonlinear wave propagation along
viscoelastic bar 343
Normal stress effect, see
Weissenberg effect

472

Subject Index

Nucleation speed 405


Nylon 66, 357
Oldroyd 8-constant model 35
Open channel (elastic) siphon 338,
342-3
Orthogonal superposition of
oscillation on steady shear flow
data 134-7
theory 144-151
comparison with data 144-151
Oscillations in melt fracture 365-8
see also Spurt melt fracture
Overshoot phenomena 124
see also Simple shear
Parallel plate device (plastometer)
80
Parallel superposition of oscillation
on steady shear flow
data 134-7
theory 137-42
amplitude effect 143-4
Peclet number 283
Ph an Thien/Tanner model 34, 71
Phenomenological constitutive
relations 14, 15
PIB, 5.39% solution in cetane 145,
150
PIB, 6.86% solution in cetane 146,
150
PIB, 8.54% solution in cetane 147,
150
PIB, a solution in tetraline 403
Planar deformations 19
Planar extension, see Pure shear
Poiseuille flows 232--46
steady 233--4
unsteady 236--42
hardening phenomena 234-6
pulsatile pressure drop 242-5
Poisson-bracket formalism 54, 37
Polariscope 111
Polyacrylamide, PAM, 0.25%
aqueous solution 403
Polybutadiene, hydrogenated 223
Polybutadiene, PB 247, 359, 368,
372, 375, 396-98, 400, 407

Polydimethyl siloxane, PMDS 247,


364, 378, 379, 380
Polyesters 401
Polyethylene oxide, PEO, a water
solution 338, 339, 342, 343, 401
Polyethylene, PE 357, 359
Polyisobutylene, PIB 356, 401, 404
Polyisobutylene, PIB, P-20 119,
120, 143, 174, 175, 177, 179-82,
184, 185-7, 207, 208, 263, 278,
279, 296, 316, 323, 324, 332-5
Polyisobutylene, PIB, Vistanex LMMH (Enjay) 236, 241, 242, 257,
258
Polyisoprene, PI 247, 396, 398, 400,
407
Polymer chains 3
Polymer degradation and oxidation
thermal 406
mechanical 407
Polymeric liquids 1
Polymethyl methacrylate, PMMA
247, 357, 358
Polypropylene, PP 302--4, 358, 359,
368, 405
Polystyrene TC 3-30 Shell 292,
293-5, 359
Polystyrene, PS 119, 120, 133, 134,
247, 263, 356, 358, 359
Polytetrafluoroethylene, PTFE 357,
358, 359, 368
Polyvinyl butyral, PVB 358
Polyvinyl chloride, PVC 358
Polyvinylidene fluoride, PVF 359
Potential function of tensor
argument 419
Power-like elastic potential
calculating shear deformations
161-2
Power law 300, 316
Polarizability 11 0
Polymer blends
HDPE/LDPE 361
LLDPE/LDPE 361, 363
Pressure-controlled flow 284
Pressure controlled capillary
rheometer 364
Pressure gradient (drop) 226, 238

Subject Index

Principal axes of second-order


tensor 416
Principal elastic strains 17, 21
Principal values of second order
tensor 416
PS, 32.6% solution in
polychlorinated biphenyl
(Arochlor 1248) 119, 120, 121,
127, 128, 129
PS, 57% solution in decaline 156
Pure shear 353
Quick thermal adaptation 291
Rate-controlled flow 113, 284
Rational mechanics 39
Recoverable strains 1, 14
Rectilinear steady flows
general theory 224-7
predictions of Maxwell-like
models 227-32
existence 228-9
weak secondary flows 229-30
strong secondary flows 230-32
Refractive index 109
Refractive index tensor 110
Relative viscoelastic kinematics 39
Relaxation
after step shear strain 118, 131-2
after cessation of steady shear
flow 118, 120, 130-31
Relaxation function 26
Relaxation mode 3
Relaxation spectrum 27
Relaxation time 2, 4
Reptation theory 50-51
Residual length 321
Restructuring velocity in flow with
slip 394
Retardation, see Shear strain
recovery
Reynolds number 246, 381
Rheological parameters
evaluation 28-30, 122-4
linear 28
nonlinear 29-30
temperature dependence 30
Rheology of filled polymers 408

473

Rheology of liquid crystalline


polymers 409
Rheology of viscoelastic liquids
experimental methods 78
Rheometrical flows 117
Rheooptics, see Flow birefringence
'Ripple effect', see Extrudate
appearance
Rivlin-Ericksen tensor 42
Rotary viscometer
rational design 315
Rotating shaft sealing 274, 280-82
Rouse model 48
normal relaxation modes 48
Rupture of polymer melts in
extensional flow 206, 396-400
brittle rupture 397
rubber-like rupture 398
neck formation 399
'Sausage link', see Extrudate
appearance
Scalar functions of tensors 418-19
Schallamach's waves 385, 388
'Second-order' liquid 42
Second-rank tensor 413
Second Law of Thermodynamics
14-16, 68, 262, 425
Second normal stress difference 79,
116, 128
Self-consistent approach 53
Self-oscillations
in friction of elastomers 387-8
in polymer flows 387-93
cycle 390
SEPARAN AP30, aqueous
solutions 401
'Sharkskin' 357, 362, 361, 370-72
see also Melt flow instabilities
Shear rate 79, 115
Shear strain recovery 155-7
time 31
Shear stress 79, 116, 277
Shearing flows 261
Shock waves 346, 347
see also Nonlinear wave
propagation along viscoelastic
bar

474

Subject Index

Short memory assumption 42


see also Rational mechanics
Simple liquids 41
thermodynamics 41
Simple shear 79
see also Couette flows
Sliding friction of crosslinked
rubbers 383-8
characteristics 384
Slip velocity 375-6
Slope change in flow curve 360-63
Smith envelope, see Failure
envelope
Specific entropy 424
Specific free energy 424
Spiral flow, see Helical flow
Spring-and-dashpot model 25, 26
Spurt melt fracture 357, 368, 372-9
compressibility effects 376, 378
'Spurt flow', see Spurt melt fracture
Start-up shear flow
under given shear rate 124-9
under given shear stress 152--4
State variables 424
States of deformation 7
initial 6, 7
actual 6, 7
unloaded 6, 7
Static friction 385, 388
Statistical mechanics
rubber elasticity 43-5
dilute polymer solutions 45-9
polymer melts and concentrated
solutions 49-53
Statistical nonlinearity 50
Steady state shear flow 129, 130
Stick-slip 385
Stick-slip melt fracture, see Spurt
melt fracture
Strain-stress relation 11
linear Hookean 11
Finger 11
Strain compatibility 9
Strain gradient tensor 6, 413-14
total 6
recoverable 6
irreversible 6
Strain history

piecewise smooth 69
Strain measure 416
Strain rate tensors 7, 420
total 7
reversible 7
irreversible 7, 14, 16
Strain tensors, left and right 414
Strands
statistics 49
slippage 50
Strength of polymer melts 396
Stress-elastic strain relation 264
Stress-optical constant 110, 240, 258
Stress-optical law 110, 112
Stress controlled flow 114
Stress relaxation 124-9
Stress tensor 10
total 10
equilibrium 10, 36
Striking a viscoelastic bar 348
Strong discontinuities, see Shock
waves
Strong hyperbolicity 347
Sub-Networks 2
Supercritical non-isothermal flow
regime 272
'Superextrusion' 357, 359-60
see also Melt flow instabilities
Supressing retardation 296
Surface tension forces 281
Swelling ratio 327
see also Extrudate swell
Symmetry of stress tensor 423
Tapering effects 246
Teflon capillaries 375
Temperature shift factor 291
Tensor invariants 416-17
Tensorial time derivatives
upper-convected 7, 57
co-rotational 7, 34, 56
lower-convected 8, 57, 422
Thermal balance equation 283--4
Thermal entrance effects 282
Thermal flow hysteresis 286, 290
Thermal flux 423
Thermo-physical constants 262
Thermodynamic (in)stability 12, 16

Subject Index

Thermodynamic fluxes 14, 55-6, 58


Thermodynamic forces 14, 55-6, 58
Thermodynamic potential 55
Thermodynamic stability 21, 64
Thermorheological equations 260,
269, 345
Thixotropy 126
Time-strain separability 67
Time-temperature superposition
260, 264, 266, 267
Total energy 423
Total energy flux 423
Transient shear, see Start-up shear
flow
Trouton viscosity 173
Uniaxial extension
schemes of devices 100-102
methodological aspects 102-3
defects and errors in
measurements 103-4
Unloading 3, 5-6
'Unloading in small' principle 6, 39
see also Eckart's concept of
'variable elastic state'
Upper-convected Maxwell model
47,70
Velocity gradient tensor 7, 420
total 7
actual 7
reversible 7
Viscoelastic criterion for melt
fracture 379-80

475

Viscoelastic kinematics 5-10


relation 7
logical imperfections 10
Visco metric coordinate system 67,
225, 268, 311
Viscometric flows 299
Viscosity 2, 4
Volumetric relations 8
Vorticity tensor 7, 420
total 7
reversible 7
irreversible 7
Vulcanized rubbers, see Crosslinked
rubbers
'Waviness' 357, see Extrudate
appearance
Weak discontinuities 346
see also Nonlinear waves
propagation along a
viscoelastic bar
Weakly non-uniform extension
kinematics 106
schemes for drawing 107-9
spinnability 107
observations 109
relation to uniform extension
319-25
Weissenberg effect 114, 274
White-Metzner model 134
WLF approach 265
Ziegler variational principle 19

Potrebbero piacerti anche