You are on page 1of 242

DRAG REDUCTION OF TURBULENT FLOWS BY ADDITIVES

FLUID MECHANICS AND ITS APPLICATIONS


Volume 32

Series Editor: R. MOREAU


MADYLAM
Ecole Nationale Superieure d' Hydraulique de Grenoble
Bolte Postale 95
38402 Saint Martin d' Heres Cedex, France

Aims and Scope of the Series


The purpose of this series is to focus on subjects in which fluid mechanics plays a
fundamental role.
As well as the more traditional applications of aeronautics, hydraulics, heat and
mass transfer etc., books will be published dealing with topics which are currently
in a state of rapid development, such as turbulence, suspensions and multiphase
fluids, super and hypersonic flows and numerical modelling techniques.
It is a widely held view that it is the interdisciplinary subjects that will receive
intense scientific attention, bringing them to the forefront of technological advance-
ment. Fluids have the ability to transport matter and its properties as well as
transmit force, therefore fluid mechanics is a subject that is particulary open to
cross fertilisation with other sciences and disciplines of engineering. The subject of
fluid mechanics will be highly relevant in domains such as chemical, metallurgical,
biological and ecological engineering. This series is particularly open to such new
multidisciplinary domains.
The median level of presentation is the first year graduate student. Some texts are
monographs defining the current state of a field; others are accessible to final year
undergraduates; but essentially the emphasis is on readability and clarity.

For a list of related mechanics titles, see final pages.


Drag Reduction
of Turbulent Flows
by Additives

by

A.GYR
Institute of Hydromechanics
and Water Resources Management,
Swiss Federal Institute ofTechnology,
Zurich, Switzerland
and
H.-W. BEWERSDORFF
Department of Chemistry and
Chemical Engineering,
Fachhochschule Lausitz,
Senftenberg, Germany

SPRINGER-SCIENCE+BUSINESS MEDIA, B.V.


Library of Congress Cataloging-in-Publication Data

Gyr. Albert.
Drag reduction of turbulent flows by addltives / by A. Gyr and H.
-W. Bewersdorff.
p. cm. -- (Fluid mechanics and its appllcatlOns v. 32)
Includes blbl iographlcal references and lndex.
ISBN 978-90-481-4555-3 ISBN 978-94-017-1295-8 (eBook)
DOI 10.1007/978-94-017-1295-8
1. Turbulence. 2. Frictional resistance (Hydrodynamlcs)
1. Bewersdorff. H.-W. II. Title. III. Serles.
TA357.5.T87G97 1995
620.1 '064--dc20 95-12572

Printed an acid-free paper

Ali Rights Reserved


1995 by Springer Science+Business Media Dordrecht
Originally pub1ished by K1uwer Academic Publishers in 1995
Softcover reprint of the hardcover 1st edition 1995
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.
Table of Contents

I INTRODUCTION 1
1.1 The phenomenon of drag reduction 1
1.1.1 The discovery of a new effect 1
1.1.2 A review of laminar behaviour of dilute drag reducing solutions 3
1.1.2.1 A Couette flow experiment 3
1.1.2.2 The Weisenberg experiment 4
1.1.2.3 A tearing test 4
1.1.2.4 Other phenomena of laminar flows 5
1.1.3 The phenomena in turbulent flows 5
1.1.3.1 The onset phenomenon 6
1.1.3.2 Virk's maximum drag reduction asymptote 7
1.1.3.3 Changes in the structure of turbulence 7
1.1.3.4 Degradation 7
1.2 Definition 8
1.3 Representation of data under drag reducing flow conditions 9
1.3.1 The friction factors 10
1.3.2 The mean velocity profiles 13
1.4 A short survey ofreviews on the subject 14
1.5 A short survey of the contents of this book 15
II PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND
SUSPENSIONS, SURFACTANTS IN SOLUTIONS; 17
CHARACTERIZATION OF FIBRES

2.1 Some general remarks 17


2.2 Polymer molecules in solution and suspension 17
2.2.1 The dilute solutions 17
2.2.2 Models of dissolved polymer molecules 18
2.2.3 Degradation 24
2.2.4 Molecular weight and molecular configurations 26
2.3 Surfactant solutions 27
2.4 Characterization of fibres 30

ill RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS, AND OF


FIBRE SUSPENSIONS 33

3.1 Introduction, rheological terminology 33


3.2 Rheometry 40
3.2.1 Shear viscometers 40
3.2.2 Elongational Viscometers 45
3.3 Rheology of dilute polymer solutions 46
3.3.1 Constitutive equations 55
3.4 Rheology of semi-dilute polymer solutions 57
3.5 Rheology of surfactant solutions 60
3.6 Suspension rheology of solid particles 65
IV DRAG REDUCTION AND TURBULENCE 69

4.1 Introductionary notes 69


4.1.1 Equations based on conservation laws 70
4.1.2. Reynolds decomposition 76
4.1.3. The flow equations in turbulent decomposition 77
vi TABLE OF CONTENTS

4.2 Additional statistical means of turbulence 82


4.2.l.Correlation as definitions of scales 82
4.2.2 Spectral description of turbulence 88
4.2.3 The concept of coherent structures 92
4.3 Fundamentals used to explain drag reduction 94
4.3.1 A rheological explanation 94
4.3.2 Explanations based on change in the energy transfer in the turbulent flow 96
4.3.3 Explanations by scaling arguments 97
V DRAG REDUCTION IN POLYMER SOLUTIONS 101
5.1 Integral aspects or gross flow behaviour 101
5.1.1 The characteristic parameters 101
5.1.1.1 Dilute solutions 101
5.1.1.2 Semi-dilute solutions 104
5.1.2 Up-scaling problem 106
5.2 The turbulent velocity field 108
5.2.1 Turbulent structures in statistical decomposition 109
5.2.1.1 Turbulence intensity 109
5.2.1.2 Reynolds stresses 111
5.2.1.3 Joint probability density function and higher moments 113
5.2.1.4 The kinetic energy distribution, production and dissipation 117
5.2.2 Turbulent structures viewed as "coherent structures" 120
5.2.2.1 The structures detected by velocity measurements 120
5.2.2.2 Topological aspects of drag reduction 122
5.2.2.3 Visualization of structures 126
5.3 Heterogeneous drag reduction 126
5.4 Drag reduction under rough wall conditions 130
5.5 Theories of drag reduction, or discussion of hypotheses 133
5.5.1 The introduction of new length and time scales by the additives 134
5.5.2 Changes in the burst process, an introduced stability mechanism 143
5.5.2.1 Landahl's instability theory 144
5.5.2.2 Stabilization of the bursting 146
5.5.3 Development of anisotropic low dissipative structures 147
5.5.4 Virk's type B drag reduction 150
5.5.5 Some concluding remarks 153
VI DRAG REDUCTION IN SURFACTANT SOLUTIONS 157
6.1 Friction Behaviour 157
6.2 Velocity profile and structure of turbulence 161
6.3 Rheological properties influencing drag reduction 170
6.4 Hypothesis on the drag reducing mechanism by surfactants 172
6.4.1 A hypothetical rheological explanation 172
6.4.2 A fibre-like explanation 17 4
VII DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 175
7.1 Drag reduction by non-fibrous suspensions 175
7.2 Drag reduction by medium size fibrous suspensions 17 8
7.3 Drag reduction by long fibrous suspensions 183
7.4 Additional effects 190
VillAPPLICATIONS 191

8.1 Introductory remarks 191


8.1.1 General economic considerations 191
TABLE OF CONTENTS vii

8.1.2 Environmental considerations 192


8.1.3 Optimimization of the selection process for the additives 193
8.1.4 Systematics 193
8.2 Drag reduction in crude-oil pipelines 194
8.3 Drag Reduction in Sewer Flows 195
8.4 Central Heating Systems 200
8.5 Ships and submerged bodies 202
8.6 Separation and Hydrofoils 203
8.7 Hydrotransport 204
8.8 Cavitation protection and noise suppression 205
8.9 Stabilization of sediment transport 206
8.10 Biological applications 207
8.11 Irrigation systems 208
8.12 Possible future apllications 209
8.12.1 Improvement of already known applications 209
8.12.2.1 Combinations of different drag reducing systems 210
8.12.2.2 Heat exchanger 210
8.12.2.3 Drag reducing fluids in rotating flows or in flows with curved walls 211
8.12.2.4 Jet and film flows 211
8.12.2.5 Drops and bubbles 212
8.12.2.6 Laminar - turbulent transition 213
8.12.2.7 Manipulation of separation 213
8.12.2.8 Entry and exit flows 214
8.12.2.9 Exotic applications 214
8.12.2.10 Application in recreation and sports 214
8.12.2.11 New methods of adding 215
8.13 Preparation of polymer and surfactant solutions and fibre suspensions 215
IXREFERENCES 219
XINDEX 231
Preface

In the present series several books have already been published on drag reduction in
turbulent flows. These books deal with manipulations of turbulent flows by passive means,
especially by attaching special roughness elements to the wall, e.g. riblet sheets. However,
what has been missing is a description of the drag reduction in turbulent flows achieved by
additives, whose efficiency is even higher than that produced by passive means.
Drag reduction by additives for fibres was discovered 60 years ago and 50 years ago for
polymers and surfactants. Since the discovery of the effect there has been an up and down
of the basic research and the testing of possible applications. This varying interest in the
effect resulted from an overestimation of the possibilities of application and the resulting
disappointments. Nevertheless, the effect is still of importance for two reasons:

1. In many cases it is easier to manipulate flows by additives than by passive means,


and the potential of applications is considerable, as we will show.

2. The effect is of importance for the research on turbulence, because every drastic
manipulation allows conclusiom ,. g the basic elements of a process.

This book covers both aspects, because ction of the basic research and the applied
research make worth wile contributions. Thus it is the aim of the book to present the state of
the art with respect to both aspects in order to motivate researchers doing basic research as
well as doing applied research in the future. Consequently, this defines the readers. There
is the researcher in the field of fluid mechanics to whom we will demonstrate in which
respect research on this effect is important for the understanding of turbulent flows. On the
other hand, there is the engineer who is looking for new, creative methods to develop
energy-saving processes in which friction losses are of importance. Furthermore, we
intend to give impulses as to how the knowledge of the effect can be useful for other areas
of research in which fluid mechanics play a minor role, e.g. methods of irrigation in
agriculture. Thus besides the basic research and the application of the effect the book is also
interesting for people working in other fields and willing to integrate new ideas for their
purposes.
Drag reduction by additives cannot be reduced to one effect and its explanation.
Surprisingly, there are several effects which are very similar and result in the same
outcome. Thus a broad description of the effect is essential.

ix
X PREFACE

The book is divided into three parts: In the first part basic knowledge is more broadly
presented than necessary for the understanding of the following parts. This is done to
enable the reader to develop and formulate his own thoughts. The second part of the book
is the core part. In this part drag reduction by addition of polymers, surfactants, and fibres
is presented. Relevant experimental results as well as theoretical attempts or hypotheses to
explain the effect are presented. In the third part the potential of applications is discussed. It
ends with a subchapter, in which several possible future applications are speculated at.
The book is based on the research results of the last 60 years. Consequently, many results
and explanations were found several years ago. Thus the subject will not only be elucidated
in view of the most recent publications in this field. Also older results, ideas, and
explanations are included in order to gain a complete insight. This was taken into account
by referring to older, original publications and resulted in an extensive list of references.
Finally, we hope that this book can help to point out how interesting the effect of drag
reduction by additives is.

ZUrich, Senftenberg January 1 9 9 5


Acknowledgements

We thank many friends who helped us with their advice during the preparation of the book.
Our special thanks goes to Professor Dr. T. Dracos, who enabled the collaboration of the
two authors. Furthermore, we thank Professor Dr. A. Tsinober for his friendly, critical
comments and discussions of parts of the book, and Mrs. A. Vyskocil for improving the
English.
Last but not least, we thank Edith and Margret, for their patience and support during the
preparation of the book.

xi
I INTRODUCTION

1.1 The phenomenon of drag reduction

1.1.1 The discovery of a new effect

When Toms (1949) discovered that minute amounts of added long-chain polymer
molecules could reduce the pressure losses in a turbulent pipe flow, this was not the first
time that such an effect by additives was observed. It was already known that fibre
suspensions, mainly fibres used in the paper making process, showed the same effect in
concentrations, however, that were two orders of magnitude higher than in Toms'
experiment (for a review of these earlier papers see Radin et al. 1975). Toms however,
showed for the fust time that minute amounts, 5-10 ppm per weight, of polymers could
have a really dramatic effect. In other words, the drag reducing effect became a promising
candidate for partial solutions of the problem of energy saving.
Toms' discovery did not remain the only one in this field. In the meantime it was
recognized that drag reduction is a much wider spread phenomenon than originally
thought. Drag reduction by a new variety of additives as well as by diverse other
manipulations, such as by strong magnetic fields or special wall roughnesses, were
found. However, the additives remained the most efficient and we will restrict ourselves
in this book to this category of drag reducers only. Review articles of the most common
drag reducing effects are found in the books by Hough (1980) and by Bushnell and
Hefner (1990). As for us, we have restricted ourselves to a table giving a survey of the
different effects and their efficiency (Table 1.1.)
The observation of a reduced pressure loss by minute concentrations of additives is one
thing, whereas an explanation of it is another, and it is the explanation which is needed to
adequately manipulate the flow in a given application. This will be the main goal of this
book.
Beside the great impact of this effect on the basic studies of turbulence it also has a
remarkable contribution in solving environmental problems or in ma:king industrial
processes more cost-effective. For example, in pipeline systems in which, due to the
installed pumps, the pressure drop is constant, drag reduction results in an increase in
discharge. This type of application is used in the Trans-Alaska-Pipeline, the most
successful application of drag reduction achieved by adding polymers. The adding of
polymers to the crude oil in this pipeline started in 1979. The desired discharge of two
million barrels per day could be obtained without the construction of two additional
pumping stations planned, Motier & Prilutski (1984 ). Further, currently used
applications of drag reduction are the addition of surfactants to closed flow circuits, e.g.
2 CHAPTER I

in district heating systems in order to reduce the pumping costs, or the dosing of
polymers to sewers in order to increase their discharge capacity during heavy rainfalls.
These and other applications will be discussed in chapter 8.

Table 1.1: Diverse dra reducin effects


Active Drag Reduction Magnitude of drag reduction by additives or by body forces

Dust in gas

Sand suspensions

Clay suspensions ~ . ~.. ..

Injection of gas (bubbles) ...


Flakes, flocks, algae and biological
molecules

Paper pulp, and similar material . .......... .

Fibres (asbestos)

Surfactants anionic

cationic

nonionic

Dilute polymer solutions depending on concentration, length and kind

stiff molecules

elastic molecules

Heterogeneous DR

continued thread

breaking into pieces


depending on concentration, relative velocity and kind

, ........
......._. ---
Combinations of polymers,
surfactants and fibres

Intelligent manipulation by
injections

MHD parallel field

transverse field
INTRODUCTION 3

Enhanced laminar stability suction, injection, heating, etc. will not be discussed here

Passive Drag Reduction geomeuical effects (special boundary and initial conditions will not be
discussed here)

LEBU

Rib lets

Special Roughness

Compliant walls passive

active

controlled active

Fluid Films

Convex Curvature
not quantified

% experimentally always achieved

% experimentally not always achieved

1.1.2 A review of laminar behaviour of dilute drag reducing solutions

In effect, the fascination of the drag reducing effect is strongly related with the dilution of
the solution, since it seems incredible that such small amounts of additives are capable of
changing the energy dissipation of a turbulent flow so drastically. This is only
understandable if the additives are capable of changing the flow structure, and this raised
the question whether the additives in such a dilute form are not only changing the
dynamics of the flow but also changing the fluid properties in comparison with the
solvent. Such properties can best be studied in laminar flows; therefore a series of simple
experiments are given to get a better understanding of the properties of the solutions.

1.1.2.1 A Couette flow experiment

It is common practice to measure the viscosity 11 of a fluid by exposing it to a laminar


Couette flow. That means the fluid is filled into a narrow gap between two concentric
cylinders. The inner cylinder has the radius R1; the outer, a radius R2. One of these
4 CHAPTER!

cylinders can be rotated with respect to the other. The fluid in the gap exhibits a shear
stress which can be measured via the torque on the cylinder in motion.
If one assumes that the viscosity is the proportionality factor between the shear, 't, and
the velocity gradient over the gap, its value can be evaluated by

(1.1)

with U<p the circumferential velocity of the fluid. Fluids whose properties under the
aforementioned deformations can be described by the single material constant ll are called
Newtonian fluids.
Evaluated in such a way ll for a dilute solution or suspension was found to be only a few
percent higher than that of the solvent. This means the dilute solution behaves under
Couette flow condition as a Newtonian fluid with a slightly increased viscosity. This was
not an unexpected result.

1.1.2.2 The Weisenberg experiment

If a concentrically positioned cylindrical rod is rotated in a fluid with a free surface filed
up in a cylindrical vessel, a water like fluid is brought into motion so that at the centre of
rotation one observes a depletion in the free surface which is due to centrifugal forces.
Not so for the dilute drag reducing solutions. The solutions start to climb on the wall of
the rotating rod. The solution behaves in a non-Newtonian manner. The effect is known
for concentrated polymer solutions or melts and is called the rod-climbing or Weisenberg
effect. For such dilute solutions this was unexpected.

1.1.2.3 A tearing test

If one pulls out a glass rod from a beaker filled with water, a certain small amount of
water is removed with the rod and drops afterwards down into the beaker. The same
experiment made with a dilute drag reducing solution shows a new phenomenon. In the
gap between the rod and the fluid in the beaker a thin filament of fluid is created. When
this filament breaks apart, it does not form drops; it snaps like an elastic band. Again the
dilute polymer solution does not exhibit Newtonian behaviour, and since it has the
mentioned elastic character, it is called a visco-elastic fluid.
INTRODUCTION 5

1.1.2.4 Other phenomena of laminar flows

Since the drag reducing solutions do not behave as Newtonian fluids, some additional
experiments have been performed to gain better understand the differences between the
solution and the solvent. It was found that when these solutions are exposed to a laminar
contraction flow, like that encountered in tube entrance flows, the pressure drop is
increased. Porous media comprise statistically distributed locations of contractions, and
indeed the pressure drop increased drastically for the solutions, see e.g. Durst et al.
(1980). This additional effect is riow widely used, for example for a better yield in oil
production.

1.1.3 The phenomena in turbulent flows

What we know is that certain dilute solutions or suspensions are capable of reducing the
pressure drop of a turbulent pipe flow compared with the flow of the Newtonian solvent.
Usually additives enhance the drag; therefore one has to evaluate which additives and
under what conditions these additives are drag reducing.
First, it was found that only additives with an extreme aspect ratio and a rather high
flexibility are effective and these additives are most efficient in very low concentrations.
This means that for any type of drag reducing additives the effect is largest at a specific
rather low concentration. If the concentration is increased, the effect decreases and
finally, at even higher concentrations, disappears. It is thought that this concentration
dependence is due to some kind of particle-flow and the particle-particle intemction. If the
latter becomes dominant, no drag reducing effect can be observed anymore.
From the investigations of the laminar flow behaviour we know that if the drag reducing
solutions are exposed to a simple shear flow, they exhibit Newtonian behaviour.
However, as the solutions are exposed to a strong strain field, the flow resistance
increases drastically, see chapter 3. One hypothesis, therefore, is that the effect is due to a
stretching and alignment of the long additives, which produces at least a local
"reinforcement" of the fluid. In this configumtion the particles do not interact directly
with each other. But even if the additives in such low concentration are stretched and
aligned, are they capable of influencing the flow field? A crude estimate will give an idea
what polymers in solution would look like.
In the estimate we assume a species of polymer molecules with a molecular weight of
lo4kg/mol in a concentmtion of lOppm. If the fluid is at rest, the molecules form spongy
balls with a mdius,a"' 1.8I0-7m and the mean distance between the centres of mass of
the molecules is r"' 1.2I0-6m, Fig. l.la, and therefore r/2a"' 3.3. If a strain field is
6 CHAPTER I

applied, the balls start to deform into ellipsoids, which in their most extended fonn
become needle like with a long axis b"" 1Q-5m and short axis c"" 2.7I0-8m. If these
"needles" are aligned, see Fig. l.lb, the distance, d, between them in tranversal direction
is then d""410-7 and d/b"" 410-2. In other words, although the molecules do not interact
with one another, the system cannot be called a dilute one anymore and the effect must be
due to an interaction of the additives with the structures of the turbulent flow. Although
the mechanism for how the state (b) is achieved is not explained here, this configuration
will be helpful in understanding the models explaining the effect in chapter 5.

(a) (b)

~
-5
c""10 m
...

25000:1 2 500: 1

Figure 1.1 Configuration of long polymer molecule of M""l04kg/mol in water at a


concentration of lOppm. (a) in the fluid at rest and (b) in a pure strain flow at
maximum elongation.

1.1.3.1 The onset phenomenon

The onset phenomenon is restricted to polymer molecules as additives and is mentioned


already in this general context since it provides evidence for the above hypothesis. The
drag reducing effect starts at Reynolds numbers; or wall shear stresses respectively,
larger than some threshold which depends on the concentration and the kind of additives.
Below this threshold the flow behaves like a Newtonian fluid.
Due to the hypothesis the molecules must be long enough and be aligned. To satisfy this
criterion, the polymer must encounter a strain field strong enough to pull them out and to
align them. In a turbulent flow field the locations of high strain are distributed in space
and time. In the frame of this idea the onset indicates that enough regions are present
where the molecules can be stretched and oriented and act by a not yet understood
mechanism as drag reducer. It will therefore be one of the goals of every theory to predict
the onset due to the characteristic parameters of the additives and the turbulent flow field.
INTRODUCTION 7

1.1.3.2 Virk's maximum drag reduction asymptote

From the arguments in the above paragraph, the drag reduction must be limited by a
configuration in which all additives are elongated and aligned, in this final state With the
mean flow, and indeed any dilute polymer solution has a state of maximum drag
reduction. Virk found additionally the empirical law that all dilute polymer systems have
an asymptotic behaviour: Maximum drag reduction is limited by the so-called Virk's
MDR asymptote, which appears in the velocity profile as well as in the friction diagram.
More details are given in chapter 1.3.1 and 1.3.2.
Up to now it was believed that this limit is not so much dependent on the additives but
must be a feature of the turbulent flow, a hypothesis which makes the drag reducing
effect so extremely interesting from the view-point of turbulent research. However,
several newer experimental results, mainly with drag reducing surfactant solutions, show
that this assumption probably does not hold in its full generality, and the effect is
therefore more complex insofar as different additives act differently. Whether a specific
additive has its own MDR asymptote or not is still an open question.

1.1.3.3 Changes in the structure of turbulence

Changes in the structure of turbulence do not follow from the overall flow characteristics
as pressure drop or the velocity profile and shall be mentioned here in a
phenomenological sense to help the reader follow some of the speculations which will be
mentioned before the results themselves are presented in chapters 5 to 7. The change in
the turbulent structures discern with the different type of additives used. However they all
have in common that they have an enhanced anisotropy of the field of turbulent velocity
fluctuations. The velocity fluctuations in flow direction are normally enhanced, whereas
the ones perpendicular to this direction are reduced. Additionally, the correlation between
these fluctuations responsible for the Reynolds shear stress is strongly reduced, too.

1.1.3.4 Degradation

The drag reduction by additives terminates at a certain Reynolds number. There are two
possible reasons. (1) Some of the additives are destroyed or their specific configuration
necessary to produce drag reduction breaks down and/or (2) the particles start to get
misaligned. In case (1) degraded additives can be identified and the efficiency of the
additives decreases. In the second case the particles get disordered and with the disorder
in orientation drag reduction disappears and the drag even increases. If a breakdown of
8 CHAPTER I

the effect is recognized in a system in a maximum drag reducing state, one also uses the
name retro-onset for the critical value of wall shear stress at which it starts.

1.2 Definition

Drag reduction was discovered as a reduction in the pressure drop of the flow of a
turbulent pipe flow at the same flow rate due to additives. In this sense the terminology
was used in this text. Equivalently it can be fonnulated that the additives reduce the wall
shear stress, 'tw, or the skin friction. It is this property by which Lumley (1969)
proposed a definition of drag reduction, which is used throughout this book:
Drag reduction is the reduction of skin friction in turbulent
flow below that of the solvent.
This implies that the flow of the solution or suspension has to be turbulent, and that the
skin friction is lower than that of the Newtonian fluid having the same viscosity as the
solution at the wall shear stress in question. As we have seen, the drag reducing solutions
generally are non-Newtonian. The viscosity is not a constant anymore and this will be
one of the main problems in describing drag reduction phenomena. The implications of
this fact will be discussed later on in some detail.
In the frame of this definition, drag reduction, DR, is given by thefrictionfactorratio or
by the ratio of the wall shear stress , 'tw, for the solution with the suffix s and the
Newtonian solvent with the suffix N:

DR= 'tws
'twN (1.2)

The wall shear stress and Fanning's friction factor f, for pipes are related by

f=~
P-2
-UQ
2 (1.3)
where UQ is the bulk velocity given by

-UQ=-
QA
p (1.4)
with Q being the discharge or flow rate and A the cross-section of the pipe.
Therefore eq. (1.2) can also be written as
INTRODUCTION 9

DR= fws
fwN (1.5)
for the same discharge condition and, Pw Ps
Most common , however, is to report on drag reduction by giving the percent drag
reduction

f
DR(%)= 1-_!._ or 1- Aps x100
fN \pN (1.6)

with L\p being the pressure drop over the considered pipe length. DR is often used for
DR(%).

1.3 Representation of data under drag reducing flow conditions

The problem is how the results of such flows shall be presented since, in the cases
mentioned, the relevant flow parameters are no longer dependent simply on the Reynolds
number but on various properties due to additives, like the geometry or the flexibility of
the additives as well as their concentrations. These properties must be considered too.
Therefore, a one parameter representation is an over simplification and needs a discussion
of the dependence on the other parameters.
In order to compare the results with different additives, we introduce here a
representation scheme in which the basic data are plotted as the velocity profile and the
skin friction. The main problem is due to the fact that the viscosity of the solvent and the
solution differ.
In the first publications on the drag reducing effect the pressure drop, L\p, in a pipe flow
with and without additives was compared, and L\p's versus Reynolds number for a given
material were plotted, or in a few cases the friction factors versus Reynolds number. The
first insights into the effect were achieved when, instead of such global information, the
local interaction was investigated by local measurements, mainly of the velocity.
According to Lumley's definition, the flows have to be compared using the same
viscosity given by the one of the solution. To simplify the comparison, the kinematic
viscosity, V=f..l/p, of the solutions, as it is found in a pure shear flow, is taken like a
Newtonian viscosity. Most information is represented in literature in this form especially
the velocity profile, and with respect to the definition the friction factor as a function of
the Reynolds number

uQD
Re=--
v (1.7)
10 CHAPTER I

with D the diameter of the pipe and v the kinematic viscosity as just mentioned.
However, in most cases the comparing measurements in the solution are made without
this correction, and therefore we find this problem of representation throughout the
literature on this subject, and only very few authors introduce a correction or use two
different Reynolds numbers in the same plot.
This problem becomes even more pronounced if one represents data in wall coordinates ,
for which the length scale is

1+ = lu't
v (1.8)
and the velocity scale is

+ u
u =-
u't (1.9)
with
1
_ ('tw
u't- )2
p ' (1.10)

since in this case also 'tw of the two fluids differs.


It has become customary to use the 'tw of the solvent keeping in mind the problem which
results when interpreting the data. In other words, the main plots of the velocity profiles
will be given as u+ versus y+ based on material properties of the solvent, and the friction
behaviour will be given as the friction factors versus the Reynolds number, Reo, or as a
Prandtl-Karman representation by f-1/2 versus Reofl/2 for Newtonian fluids. This
representation is so widely used since the functional dependence between f and Re show
up in it as straight lines. Schematic representations are given in Figs 1.1-1.3. For more
detail on the representation problem, see Bewersdorff & Berman (1988).

1.3.1 The friction factors

Fig. 1.2 shows a schematic friction factor behaviour in which the friction factor is plotted
versus the Reynolds number of a pipe flow. The skeleton of such diagrams are the curves
for a Newtonian fluid in a pipe with smooth walls. It is given by a laminar branch (1) for
which the friction factor is given by

f = _!!_
Re (1.11a)
IN1RODUCTION II

or
1 1
r-2=Rer2
16 (l.llb)
and the turbulent branch (2) by the Prandtl-von Kannan law

(1.12)

Figure 1.2 Schematic friction factor versus Reynolds number plot. 1: laminar pipe flow;
2: for turbulent pipe flows in smooth tubes; 3: Virk's maximum drag
reduction asymptote; 4, 5 and 6: three typical drag reducing behaviours for
flexible polymer, for surfactant and fibre solutions.

The two areas are separated by a gap in which the transition from a laminar to a turbulent
flow occurs due to instability processes which are not universally reproducible. The flow
can therefore be laminar or turbulent and during transition no specific curve can be given.
Virk's ultimate velocity proflle naturally has its analogy in an asymptotic friction factor

(1.13)
Its plot is called Virk's maximum drag reduction asymptote (MDR) (3). By definition the
friction factor of a drag reducing flow must have values between the Prandtl-von Kannan
law and Virk's MDR asymptote for a given Reynolds number.
The three best known drag reducing additives, long chain polymers, surfactants and long
fibres, have a characteristic friction behaviour, schematically shown in Fig. 1.2. The
polymers and the fibres have a characteristic onset of the effect; the surfactants, a retro-
12 CHAPTER!

onset. The characteristics of these three solutions will be discussed in detail in the
different chapters of this book.

2Sr-------------------------y-----------,

20
.
I
I
I I
I I
I 41
I 51
I I
IS I I
I
7 I I I
I I
I
I I
II
I
I
,II
10

5
2 2.5 3.0 3.5
log (Re f 112 )
10

Figure 1.3 Scheme of the friction-factor behaviour for dilute drag reducing polymer
solutions. 1: the Newtonian laminar regime; 2: the Prandtl-Karman curve for
Newtonian fluids in a smooth pipe; 3: Virk's maximum drag reduction
asymptote; 4, 5, 6 and 7: typical curves for polymer solutions; 4 and 5: of the
same polymer but with different concentrations; and 7: a polymer with a
lower molecular weight exhibiting a different onset point, 0; 6: a polymer
which does not reach the Prandtl-Karman curve. ois the slope increment.

It is seen from eqs. (l.llb, 1.12 & 1.13) that the frictional regime plotted, as f-1/2 versus
Resfl/2 in a logarithmic scale, must show up as straight lines in such a representation.
This is another very common description for drag reducing polymer solutions
schematically shown in Fig. 1.3. In this representation the "friction-lines" intersect with
INTRODUCTION 13

the one representing the P-K-line at an angle 8, the slope increment. Its relation with the
properties of the additives will be discussed in chapter 5.

1.3.2 The mean velocity profiles

In Fig. 1.4 we recognize the simplified universal velocity profile of a Newtonian fluid as
a full line. The three layers of the turbulent flow, the viscous sublayer, with

(1.14)

the buffer layer (5 < y+ $; 30), and the logarithmic layer given by

u+ = 2.5lny+ +5.5 (1.15)

are reduced to a two-layer model in which the buffer zone is eliminated by an


extrapolation of the two other zones to their intersection.
The velocity profiles of drag reducing flows in many cases show a peculiarity, a
logarithmic profile starting at the intersection with a steeper slope and describing the
velocity at maximum drag reduction (in Fig. 1.4 the dashed thick line). This is the
empirical law found by Virk (1971) and therefore often called Virk's ultimate profile,
mentioned in chapter 1.1.3.2 It is given by

u+ = 11. 7lny+ -17 (1.16)

The velocity profile of a non-maximum drag reducing flow intersects with Virk's limiting
profile at a certain level A and a characteristic angle a. Often a is zero and the bulk flow
of the drag reducing flow can be described as

u+ = 2.5lny+ +5.5 +i\B (1.17)

with dB corresponding to a shift of the logarithmic velocity profile of a Newtonian fluid


due to the presence of the additives. The zone in which the velocity profile follows Virk's
limiting profile is called the elastic zone or layer, since the solution exhibits in this zone
an elastic flow behaviour.
Velocity profiles of the flows with different additives are shown in the proper chapters.
14 CHAPTER I

two-layer model
wall-layer logarithmic layer

1 10 100 1000

Figure 1.4 A schematic representation of the universal velocity profiles. Explanations are
given in the text

1.4 A short survey of reviews on the subject

After the effect which dilute polymer caused to flow solutions was found by Toms
(1948), energy and energy saving in connection with the effect became a main issue. It
has been supported by the diverse institutions responsible for the funding of research. A
large amount of all sorts of results has been published in various journals. Therefore,
fairly early there already was a need for review articles to order the different results.
Lumley (1969) wrote a first condense review on the effect in the first volume of the
Annual Review of Fluid Mechanics, an article which even today features much of our
basic knowledge. White & Hemmings (1976) published a review and bibliography which
contains the abstracts of practically all papers on that matter known at the time.
In 1975 Virk published a review in which he described the fundamentals of the drag
reducing effect and used all data available at the time to formulate several empirical1aws
along with physical explanations. Practically at the same time IUTAM held the first
conference (1976) at which drag reduction was discussed in the context of new ideas in
INTRODUCTION 15

the fluid dynamics of turbulent flows. The proceedings, a special issue of Physics of
Fluids, Frenkiel et al. (1976), is, together with the Virk and Lumley papers, the main
source of knowledge at this time and contains six review papers along with many original
works.
The best survey of the progress made in the following years is a series of conference
proceedings. Among these series is one from the conferences on drag reduction initiated
by BHRA: Coles (1974), Stephens & Clarke (1976) and Sellin & Moses (1984) and
continued by IAHR, Sellin & Moses (1989). IUT AM organized two conferences which
supplemented each other. The one in Essen, Gam pert (1985), focused on the influence of
the material properties on the effect, and the one in ZUrich, Gyr (1990), on the aspects of
fluid dynamics. The conference in ZUrich, a continuation of the Washington conference,
1976, was also designed to bring together researchers in passive and active drag
reduction with the intention of discovering the common fundamentals. In these two
volumes one fmds, respectively, seven and nine review articles on special topics giving a
good state of the art.
In the series of Progress in Astronautics and Aeronautics two volumes have already
appeared with review articles on the different topics of drag reduction, Hough (1980) and
Bushnell & Hefner (1990). Additional review articles to be mentioned are the one by
Berman (1978, 1986) discussing specific results of drag reducing polymer flows and the
three reviews published by the group in Dortmund, Giesekus et al. (1981, 1985 and
1987). In the 1987 paper the authors showed that in flows of drag reducing fluids the
turbulent structure also changes in flows which are not wall-bound, e.g. jets or in mixing
layers.
Last but not least, a review article which focused on applications shall be mentioned.
Especially in part two it gives a very detailed list of the applications known at the time,
Sellin et al. ( 1982).
No claim is made to completeness, and the many references in this book clearly show that
researchers interested in the drag reducing effect are bound to survey a large number of
journals.

1.5 A short survey of the contents of this book

There are different forms for presenting the material. We have chosen what we think to
be the classical didactic form by which first the basics are provided followed by the more
specific discussion of the different effects. These and the application of the effects will be
covered in chapters five to eight.
In chapter two we outline the basic knowledge of the material properties of the additives.
16 CHAPTER!

Suitable additives are long-chain polymers and surfactants containing rod- or worm-like
micelles in their solutions. Additives can change the properties of the solutions with
respect to their flow behaviour. If the fluids are no longer Newtonian, their new
constitutive equations have to be formulated. These kinds of equations relate suitable
stress and deformation variables, or in other words, the stress - rate of strain dependence
which one needs in order to predict the deformation and flow of the fluid. This is called
their rheological behaviour. In chapter 3 an introduction to rheology is given as well as
some information on how the relevant rheological properties can be measured
Besides the technical applications of drag reduction in turbulent flows, research in this
field was also done in order to better understand turbulence by studying the changes in
structure of turbulence, which are connected with the occurrence of drag reduction. With
this aim a brief review on turbulence is given in chapter four and is meant to help in
understanding the main arguments on how the additives can probably change turbulence
to the effect that the flow is less energy consuming. For a more profound understanding
of the drag reducing mechanism this chapter would have to be much longer, and we can
imagine that some day in the future a book will appear with a title something like
"Turbulence of drag reducing fluids". Aspects of it can already be found in McCombs
(19~)0) book. The compromise we have chosen is a chapter reminding the reader of the
most essential concepts used in the description of turbulent flows. The fluid dynamicist
who is well aware of these theories may skip the text up to the chapter 4.3 where some
hypotheses of how the additives may produce drag reduction will be discussed.
Chapters 5, 6 and 7 contain specific information on the drag reducing effects by the most
common and efficient additives: dilute polymer solutions, solutions with surfactants and
suspensions of long fibres. In these chapters we discuss the interaction of the solutions
with the turbulent flow under smooth and rough wall conditions.
In the last chapter we will give a review of applications of drag reduction, including those
which are being developed and which may be developed in the future.
II PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN
SOLUTIONS AND SUSPENSIONS, SURFACTANTS IN
SOLUTIONS; CHARACTERIZATION OF FIBRES.

2.1 Some general remarks

In practice the used additives are polydisperse, and the result often difficult to inteq>ret
since it is never clear whether the effect is only the result of a subset of the used sample
or if some subsets are more important with respect to their drag reducing efficiency.
Although it is possible to produce a uniform sample of additives, it is not affordable.
Since, as we will see later on, the drag reducing efficiency of the additives varies
dramatically, it is evident that there is a need to characterize the additives by their physical
and chemical properties as well as their geometry. However, it is not the intention of this
presentation to completely characterize the additives but, rather, to discuss only properties
which are relevant for describing the interaction of the additives with a turbulent flow. In
other words, we will discuss the parameters which would also have to be considered in a
dimensional analysis, such as the molecular weight and the aspect ratio of the additives,
their extensibility and deformability. The main goal is to give information relevant for the
additive-fluid interaction on a molecular level. Such information is essential to describe
the competition between the bulk flow creating the distortion of the additives and the
entropic restoring mechanism. Therefore a discussion of the self- and particle-particle
interaction is omitted. The additives are thought to be present in a dilute concentration, a
state which has to be defined for the various kinds of additives.
Other properties of the additives such as effects of the excluded volumes, ionic charges
and knotting of the chain in polymers and flexible fibres as well as the conditions under
which some additives are produced and under which they degrade have some significance
but are only briefly mentioned.

2.2 Polymer molecules in solution and suspension

2.2.1 The dilute solutions

When we speak of drag reducing long polymer molecules, we have molecules in mind
which consist of linear sequences of N monomeric units. The number N of monomers in
the molecule can be very large, usually of the order of 105, which means that the
molecular weight is also very large, usually of the order of 106 to 107 g/mol. Usually,
these molecules are quite flexible. With the exception when some bulky groups, e.g.
17
18 CHAPTER II

phenyl groups or sugar rings, are introduced in the polymer chain the polymer becomes
more rigid.
With respect to drag reduction, mainly two kinds of water soluble polymers were
investigated, the polyethylene oxide and the polyacrylamide, Fig. 2.1

-tl- -i
H H
I I
- -c C-
I
I
H C=O
I
N,....H
'H n

Figure 2.1: Chemical structure of polyethylene oxide and polyacrylamide

These molecules are suspended or dissolved in a Newtonian solvent. In a dilute form one
can study some of the properties of single molecules, de Gennes (1990). In this context
dilute means

(2.1)

where n is the number density of the dissolved molecules, the number of molecules per
unit volume, and L the length scale of the molecules. However, in this definition a
property of the molecules is already used: its length. This quantity varies dramatically
because the geometry of the polymer chain in solution is variable. If the molecule were
stiff, and thus of rod-like shape, it would have the length of a completely stretched
molecule which is of the order of 1 ~m, whereas if it is, as Debeye & Bueche (1948)
proposed for a fluid at rest, a sphere in which the molecule is the skeleton of a porous
sponge with uniform density, its radius is called a hydrodynamic radius, Rb, which is
proportional to its radius of gyration, RG, the maximum extension of the skeleton
measured by a light scattering technique. In the case of the mentioned polymers this is of
the order of 100-200 nm. The exact geometry of the dissolved molecules is unknown.

2.2.2 Models of dissolved polymer molecules

The physical concept behind the spongy ball model is a hydrodynamical one. The
molecule, as long as it is flexible enough, has the tendency to coil up by Brownian
motion enclosing the solvent in the "cavities". Its form by symmetry arguments of the
forces acting on the molecule is spherical, but can be deformed by external asymmetric
PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 19

forces, which are present in most fluid flows. The assumption for the spongy ball model
is that the vorticity cannot penetrate into this spongy ball, and global properties such as
viscosity or sedimentation can be described by Rh.
Light scattering experiments showed that the model is qualitatively correct. Therefore this
kind of form of the dissolved molecules has mainly been accepted, with the
generalization, however, that the effective concentration in the interior of the coil is
fluctuating. The solvent can than more easily penetrate at locations where only a few
monomers are present. In this form the model of the polymer molecule is already too
complicated and a simplification is necessary.
The most encountered model is the dumb-bell model introduced by Kuhn and Kuhn
(1945). In the small elongation approximation the molecular chain in this model is
described by its elongation, r. only, which is thought to be the dominant length of the
molecule, with the separation of the two monomers at the end, Fig 2.2. Its elastic
behaviour is represented by a spring, and the sites of the frictional drag are located at the
ends of the spring, Fig. 2.2.

Figure 2.2: The schematic representation of the elastic dumb-bell approximation


for a distorted polymer in a velocity gradient, u(x,t)= Xi ()Uj(t)/oxj

For small values of r the force balance can be written as

f(-nnt11 ) + K
r:1 -_ r(ari aui(t))
- + r -
at J
- - + Kr: -_ 0
ar.J 1
(2.2)

with f being the global friction factor and K the elastic constant. In other words two
forces interact with the molecule: (1) a frictional one given by a friction which is
concentrated at the ends of the molecule and which is proportional to the velocity by
which these two centers separate and (2) a restoring one due to Brownian motion
20 CHAPTER II

equivalent to an elastic force as found in a spring proportional to the extension of the


spring. The velocity gradient tensor has to be evaluated as a function of time. in a
Lagrangian frame since the polymers are advected by the local bulk flow. In the Stokes
approximation

f = 61t'fla (2.3)

with TJ being the solvent viscosity and a the radius of a small particle, one finds

(2.4a)

and the inverse of (2.4a) is a time, 't:

f
't=-
K (2.4)

The quantity 't is called the relaxation time of the molecule and is of the order of a
millisecond in dilute polymer solutions up to seconds for polymers of high molecular
weight in very viscous solutions. In other words, for predicting the relaxation of the
molecules, for example, due to their elongation, the knowledge off and K, for example
as a function of the molecular weight, would be required. The dumb-bell model can be
interpreted as a simplification of a more physical bead-spring model. In this model the
sites exposed to fluid friction are represented by not only two but several beads. These
beads are thought to be connected by parts of the polymer chain, which itself is
considered to be frictionless. In the model the chains between beads are taken as equal in
length and are sufficiently long so that the actual separation in the 3-D space of the beads
obey a Gaussian distribution. Consequently, the chain elements can be considered as
entropic springs.
In a model in which the monomers themselves are considered as beads we would expect
that the friction of the different monomers is additive, which means for a chain with N
monomers

(2.5)

Kuhn & Rouse, with fm being the friction factor per monomer, Rouse (1953). The elastic
forces coincide with those of an ideal chain of Gaussian distributed beads.
PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 21

where kB is the Boltzmann constant and T the temperature. Ro=Rh at rest and a is the
monomer size. From eqs. (2.5 & 2.6) it follows that the relaxation time is proportional to
the square of the molecular weight

(2.7)

Since only the monomers at the surface of the coil are subject to friction forces, the
friction cannot be additive as assumed by eq. (2.5). Instead, the friction coefficient is
taken in the form

(2.8)

The relaxation time for ideal chains then follows from eq. (2.4 using 2.8 & 2.6), given
by the Zimm expression, Zimm (1956)

(2.9)

This result has to be extended for the case of chains embedded in a good solvent. Flory
(1971) introduced the excluded volume effect, by which it is expected that

(2.10)

with Rp being the Flory radius. However, eq. (2.10) cannot be directly inserted into eq.
(2.8), because the elastic constant, K, also has to be adopted to the case of a good
solvent. De Gennes (1977) introduced the free energy as a function of the elongation, r,
by scaling laws of a swollen coil. For small r/Rp Hook's law is valid and the elastic
constant becomes

(2.11)

and with eq. (2.11) the relaxation time becomes


22 CHAPTER II

(2.12)

In dilute solutions we can relate the coil volume, Rp, to the viscosity and, therefore,
rewrite the relaxation time in the mostly accepted form

(2.13)

with

~11
[11] = -
11C (2.14)

where ~11 is the increase in viscosity due to a small concentration, c, in polymers. The
relaxation time, therefore, relies on a non-trivial form of elastic energy, eq. (2.11), given
by the scaling laws.
Usually the dumb-bell approximation is used. This seems very contradictory since
intuitively the bead-spring model is a much better model of a real polymer chain. The
reason for the successful application of the dumb-bell model is that it uses only one
instead of a spectrum of relaxation times. This is acceptable since the longer relaxation
time processes are more influential than the short ones in producing viscoelasticity, and
the longest of those processes corresponds with the time scale of the dumb-bell. The
dumb-bell frequency corresponds to the frequency of the whole coil, whereas higher
deformation modes occur in the interior of the coil where they can be found by light
scattering at much higher frequencies than 1/'t. The dumb-bell approximation allows
additionally to predict non-linear rheological effects, and not just the linear ones as in the
Rouse-Zimm approach, Rouse (1953) and Zimm (1956).
These models have the unrealistic feature that the elongation, r, vanishes at rest. This can
be avoided by adding Brownian motion to the beads, which results in a distortion which
tends to a small non-zero equilibrium.
So far the approximations were restricted to small distortions of the molecules. If the
elongation becomes large, two additional important effects have to be taken into account.
(1) When the elongation becomes comparable with Ra, the hydrodynamic shielding of
parts of the chain by other parts of the chain is altered. (2) At extremely large distortions
the finite extensibility of the polymer chain becomes relevant, see Hinch (1977). Both
effects are non-linear. In order to include the shape dependence (1), the molecules have
PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 23

been modelled by elastic ellipsoids, Cerf (1951). However, the inextensible flexible
thread model, Hinch (1976), is more realistic since it also takes care of the finite
extensibility (2).
Hinch described the thread by a vector function K(s,t) of arc length, s, and time, t. To
preserve the arc length, there must be a tension, T(s,t), in the thread, and if no stiffness is
assumed, the tension is the net force transmitted over the cross-section of the thread in the
direction of the tangent, d!}ds. The preservation of the arc length in time is the condition
for the inextensibility of the thread. The tension force is balanced by viscous force, f(s).
For an element of arc this force balance can be written as

(2.15)

and shows that two orthogonal contributions are due to the tension in the thread. The
force in the direction of the tangent, d!)ds, is proportional to the increase in tension,
dT/ds. The other force is normal to d2x/ds2 and proportional to the tension, T, multiplied
by the curvature, ld2xf()s21. The main evolution equation thus becomes

(2.16)

where the three terms on the right side stand for the advected transport of the molecule in
the undisturbed flow. By the increasing tension along the molecule, it is pulled in the
direction of the local tangent, and pulled normal to itself by the orthogonal force produced
by the tension.
Eq. (2.18) is incomplete in so far as it needs an additional equation for the tension. This
is connected with the internal dynamics, i.e., the inextensibility constraint, and, as Hinch
showed, given by

() 2T _!_ ( () 2 xi ) 2 T =()xi n.. -


dx
J
(}s 2 2 (}s 2 (}s IJ dS
(2.17)

The tension is generated by the component of the strain rate, Dij

n.. =!(aui + auj)


IJ 2 dxJ dxI
(2.18)
24 CHAPTER II

in the direction of the local tangent. The tension has to vanish at the end of the thread
which defines the boundary conditions of eq. (2.17)

T=O at s=L (2.19)

With eq. (2.17) we have the additionally needed equation, and the deformation can be
calculated when merely knowing the initial shape of the thread.
For a given velocity field Eqs. (2.16) and (2.17) can be solved on a computer. Fig. 2.3 is
an example of a numerical solution for a molecule in simple shear flow starting from an
initial S-shape. The result is that flexible threads practically always rapidly straighten and
tend towards an orientation in which they are in tension. This tension then snaps the
thread straight.
In both models, for the spongy deformable molecules as well as for the flexible
inextensible threads, the molecules become oriented and elongated.

Figure 2.3 The straightening of a flexible inextensible thread in a shear flow, starting
from the S-shape. The configuration x(t,s) is plotted as a function of sat ')'t =
0, 1, 2, 3, 4. (from Hinch, 1977)

2.2.3 Degradation

One of the problems involved in the use of polymers is their degradation in strong
turbulent flows. This rupture in the polymer chain can be due to chemical reactions,
radiation, heat, bacteria or mechanical forces on the thread. In a turbulent drag reducing
flow it is probable that some of the molecules can break if the tension in the thread is high
enough. Degradation is a well observed phenomenon in drag reducing turbulent flows
and allows for estimating the forces needed to crack a molecule into two pieces. A break
in more than two pieces is almost improbable since the critical value of the stress, Tcr. is
attained at a specific location first. After the molecule broke, the tension in the molecule
PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 25

decreases immediately since the balancing friction force is length dependent and decreases
immediately with the break.
Hinch (1977) gave an instructive evaluation of the forces required to break the chain. A
fully stretched molecule experiences at its center a tension of

(2.20)

with b being the length of the monomers.


For IVyl = 105s-l corresponding to a friction velocity Ut of 0.3 ms-1, the tension Tmax
given by eq. (2.20) is 10-8 N. On the the other hand, the binding forces can be estimated
via the binding energy EB"' 5105 Ws/mol and the atomic separation 0.3 nm in such
molecules. The resulting estimate of the binding force is about 10-8 N. Thus, one would
expect the polymer to break into two in a flow with a strain rate as high as 105s-I.

(a) Mass per decade

1 2
Log mol wtlO 6

Figure 2.4 (a) Idealized flow field in a cross-slot device. (b) Molecular weight
distribution of degraded PEO, initial M= 1.4106 g/mol. Adapted from Odell et al
(1984)

Odell et al. (1984) studied polymer molecules in an extensional flow produced in a cross-
slot device, schematically shown in Fig. 2.4a. The method enabled them to (1) recognize
when the chain breaks, (2) characterize the scission products, (3) provide a measure of
the stretching force induced, and by the latter to assess the strength of the chain and the
nature of the rupture process. When it starts, the chain elongation shows up as a highly
localized birefringent line. Fig 2.4 shows the effect of overstretching a fraction of PEO
(polyethylene oxyd) of M= 1.41 06 g/mol. The initial single peak corresponding to M=
1.4106 g/mol is replaced by two peaks, one corresponding to the original M value and
26 CHAPTER II

the other to one of M= 0.7106 g/mol, which clearly shows the effect of a rupture at the
center of the extended thread as predicted by Hinch.
The behaviour of the molecules when stretched under certain flow conditions will directly
enter into the rheological description of dilute polymer solutions, see chapter 3.3. From
such considerations it follows, see chapters (5.1.1 & 5.5.1), that the onset phenomenon
can be related to the stretching process of the molecules and, therefore, we can relate this
important effect of drag reducing flows to the molecular weight, or to the relaxation time
of the single molecules, respectively.

2.2.4 Molecular weight and molecular configurations

If the polymerization of the molecules were to a high degree controlled, one would be
able to produce monodisperse or quasi monodisperse polymers. This means a sample of
the material would contain polymer of one molecular weight only. However, industrially
produced polymers are polydisperse. This means a sample is composed of fractions of
material of monodisperse molecules.
It is a well known fact that the efficiency of drag reduction strongly depends on the
polymer molecular weight. Therefore the fraction with the highest molecular weight in a
sample is the fraction which determines the drag reducing properties of the fluid. The
molecular weight of such a sample is a mean and is given as a "number" or a "weight"
mean. The number mean molecular weight, Mn, is given by

k
LNiMi
M n-- .._i=_,li--_
- k
LNi
i=l (2.21)

with Ni being the number of monomers present in the iLh fraction of the sample, which is
subdivided into k fractions, with Mi being the molecular weight of this fraction. The
weight mean is given by

k
LWjMj
- -
Mw - .!:i=~l;.,-_
k
LWj
i=l (2.22)

with Wi being the weight of the molecules in the ith fraction.


PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 27

Since the number mean eq. (2.21) reacts very sensitively to the fractions of low
molecular weight in a sample, it is very important that one declares which kind of mean
molecular weight was used in an experiment.
The molecular weight alone is not the only important parameter characterizing the
polymers. Another property of equivalent importance is the configuration of the
molecules, which can change due to the rotations of chemical bonds or by
thermodynamic motions of the molecules. In addition, in experiments with drag reducing
flows one often encounters polyelectrolytes. That means the polymer molecules have side
groups of atoms which are negative (anionic polymers) or positive (cationic polymers).
This explains why the length of these molecules can be varied by adding salt to the
solution or by changing the pH value of the solution. If these charges are neutralized, the
polyelectrolytes arrange themselves in almost classical coils. However, if they are forced
to interact with charges of the same sign, they are forced to elongate into rod-like forms,
which produce a respectable increase in drag reduction.
Each of the above additional effects results in its own time constants. It is, therefore,
clear that the polymer solutions have a whole spectrum of relaxation times. They produce
the main deviation from the Newtonian fluid behaviour.

2.3 Surfactant solutions

Aside from the classical soaps, which are the alkaline salts of higher fat acids, new
synthetically produced surfactants were developed, which also consist of polar and
unpolar molecular groups. This duality is responsible for the development of films at the
phase interfaces and the occurrence of associates of the surfactant molecules in solutions.
In other words, surfactants are low-molecular chemical compounds. Their molecules
always consist of a combination of a water-soluble (hydrophilic) and a water-insoluble
(hydrophobic) part. The hydrophobic part is a long chain alkyl rest, whereas the
hydrophilic part of the molecule is ionizable, polar, polarizable or suitable for forming
hydrogen bridges. The hydrophobic part of the molecule is in a polar medium, i.e., a
very hard soluble in water. This is compensated for by the hydrophilic head-group.
Depending on the nature of this head-group, the surfactants can be classified as nonionic,
cationic and anionic. The left part in Fig. 2.5 shows a well known drag reducing cationic
surfactant molecule as an example.
28 CHAPTER II

~~~:-(CH2 )13_.:~~ \ t r@(COOl- + Na+Br-

hydrophobic hydrophilic ~ OH J
cationic surfactant (TT AB/NaSal) or <; 4 TASal

Figure 2.5 A typical drag reducing surfactant dissolved in water with an


electrolyte. (Tetradecyltrimethylamoniumsalicylat = C14TASal)

From chemical analysis it is known that the chemical structure is closely related to the
phenomenological properties of the solutions. Above a critical concentration, CMCr, the
dissolved molecules start to form aggregates, which are called micelles. The paraffin
residues prefer, due to energy arguments, an unpolar environment, whereas the
hydrophilic head groups extend into the solvent. This is also the process by which the
surfactants produce the micelles. The inner part is formed by the hydrophobic residues,
and at the surface of the micelle one finds the hydrophilic head groups in contact with the
solvent, water. Therefore these micelles are soluble. The association of the molecules to
micelles is reversible. This means that if the surfactant solution is diluted below the
critical concentration, the micelles dissolve into monomer molecules again. The micelles
are always in thermodynamic equilibrium with the monomer molecules, and are of a size
of about 20 to 1000 monomers depending on the size and the geometry of the micelles.
Three geometrical types of micelles can be distinguished: spheres (globular), rod-like,
and discs. In dilute solutions most surfactants produce globular micelles. The structure of
these associates is defined by two opposite forces. The hydrophobic paraffin chains
avoid the energetic unfavorable contact with the solvent, water, by associating, which
corresponds to an attractive force between the alkyl chains. Therefore, the inner part of a
micelle can be thought of as being a liquid droplet composed of hydrocarbon chains that
amply fill the void of the sphere. All the hydrophilic head groups are situated at the outer
surface of the sphere and are responsible for the solubility of the micelles.
The number of monomers of a micelle, the association number, is given by the volume of
the nonpolar hydrocarbon chains and by the total volume of the micelle. The maximum
association number is reached when the total inner space is filled by the hydrocarbon
chains. This state is often already reached at moderate surfactant concentrations.
The volume of the micelles is defined by the micelle radius, which approximately
corresponds to the length of an alkyl chain in a monomer surfactant. If the concentration
is increased, the micelles can grow only until the volume is occupied by the hydrocarbon
PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 29

chains. From there on, the size of the micelles does not change anymore; the number of
micelles, however, increases.

single surfactant I globular micelles I rod-like micelles


molecules I
I
01

*
I
I 4<~
I 01I ..
..
.. . -!\..
. . . .
. .
[.....io-

01I
~-
(.l. .. ..
I
'1u:: ...

*
I @I
I
"tail" "head" I
J

c
electrical
conductivity
J.LSi

c [ppm]

Figure 2.6 A scheme of the different micelles as a function of concentration, and


the influence of a change in the micelle configuration on the electrical
conductivity of the solution

By adding some salt, i.e., adding electrolytes, or by using strongly associating counter-
ions, the electrolytic repulsion forces of the head groups can be suppressed, and the
molecules can be packed more densely and start to form rod-like or disc-like micelles, see
Fig. 2.6.
30 CHAPTER II

The length of a rod-like micelle is given by the length of the alkyl chain. The transition
concentration at which globular micelles become rod-like is called the second critical
concentration, CMCu, or more simply the transition concentration, Ct. At even higher
concentrations the micelles can form over-structures, in which the substructure micelles
can still be recognized. Globular micelles start to form a cubic network with micelles
mainly in the comers. Rod-like micelles can accumulate by flexible bindings at their end
and form three-dimensional networks.
Their mechanical properties are similar to medium-sized flexible fibres.

2.4 Characterization of fibres

Fibres are of different origins such as botanic, zoological, mineralogical, of metallurgical,


or chemical origin. By methods such as microscopy, electron microscopy, X-ray
analysis, or infrared spectroscopy, pyrolytic capillar-gas chromatography it was found
that the main property common to all fibres is the orientation of their molecules parallel to
the axis of the fibre. Based on this molecular structure, fibres have big aspect ratios, 1/d,
and their flexibility depends on the bindings of the molecules in the fibre. The aspect ratio
together with the absolute length of the fibres is the important property which defines
which hydrodynamic structures the specific fibres can interact with.
The first fibres found to be drag reducing were the fibres of paper pulp and were already
reviewed by Forgacs et al. (1953). Fibrous particles can act as drag reducers in air as
well as in liquids, and a wide variety has been tested. The characterization of fibres is in
most cases restricted to its aspect ratio- a ruther simple length scale-, and we will restrict
ourselves to the fibre suspensions in liquids.

Table 2.1

Material Size in mm aspect ratio concentration in Drag reduction


1/d %of volume in%

Nylon fibres 0.5-2.0 40-100 0.5-1.0 2-20

Rayon fibres 0.5-2.0 40-90 0.5-3.0 2-25

Cotton fibres 0.5-0.75 25-35 2.0-3.0 5-20


Asbestos fibres 0.002-0.02 1Q2-t03 0.25-2

2.0 105 25-500 ppm 20-85


PHYSICO-CHEMICAL PROPERTIES OF POLYMERS IN SOLUTIONS AND SUSPENSIONS 31

Nylon fibres, Rayon fibres, cotton fibres, and glass fibres have been investigated in
turbulent flows. A special class of fibres producing drag reduction comparable to that
found in polymer solutions are the mineral asbestos fibres. They are long and flexible and
have an aspect ratio as high as 105. Fibres which are long and very slender are the ones
of highest interest and, therefore, will be mostly discussed. In table 2.1 a schematic
representation of the length scales and the concentration required for drag reduction for
the mentioned fibres is given.

Figure 2.7 Photograph of asbestos fibres from a 300 ppm suspension. Length of bar is
lOJ.l. From Moyls & Sabersky (1978)
32 CHAPTER II

In order to gain a deeper insight into the behaviour of the fibres in suspension, several
electron photomicrographs were taken. The photographs were prepared by taking a drop
of the suspension and placing it on the grid holder of the microscope. The holder with the
samples was then placed in a chamber which was evacuated. The solvent, in this case
water, was evaporated and the residual fibres photographed. An example of such a
photography is shown in Fig. 2.7. With this technique one probably gets a useful picture
of what the distribution of fibres may look like, but this is the extent of its usefulness
since the water evaporates between the fibres. Fig. 2.7 shows how much the diameter of
the fibres can vary and that they form a rather dense mesh in which the motion of each
fibre is certainly influenced by its configuration in the mesh.
Fibres with an aspect ratio greater than 25-30 always produce drag reduction if the
concentration is sufficiently high. Since the length scale is so important, the relative
dispersing abilities of fluids of different viscosities apparently affect the drag reducing
character of suspensions more than viscosity; therefore, dispersion is a very important
parameter too.
III RHEOLOGY OF POLYMER AND SURFACTANT
SOLUTIONS, AND OF FIBRE SUSPENSIONS

3 .1 Introduction, rheological terminology

In order to understand the rheological behaviour of different materials let us consider the
following idealized experiment. A deformable fluid is found in the intervening space
between two parallel plates which have indefinite dimensions in the spanwise direction, see
Figure 3.1.

F ..

y
X

Figure 3.1 Simple shear between two parallel plates

The material does not slip at the two plates. If the lower plate is kept fixed and the upper
plate is moved by a force, E. to a distance, x, from its original position, the material in the
intervening space between the two plates will be deformed. The resulting shear in planes
parallel to the moving plate is called simple shear. In general, the displacement of the upper
plate is a function of time.
The degree of shear is given by the quotient

x dx
y=-=-=tga
h dy (3.1)

which is called strain. This common abbreviation of shear strain gives rise to some
misunderstanding. In general, strain is a relative deformation, relative to a reference
configuration of length, area or volume, and therefore also called relative deformation.
The force, E. needed to achieve a certain strain is proportional to the area, A, of the two
plates responsible for the transmission of this force into the fluid. Thus, the quotient of
these two quantities is
33
34 CHAPTER III

F
't = -
A (3.2)

It is called shear stress.


For the case of simple shear considered a material law - regardless of whether the
deformable fluid is a rigid body or a fluid - can be formulated as a relationship between the
shear stress, t, and the (shear) strain, y. In general, the time dependent behaviour of these
properties has to be considered in more detail or, more accurately, the deformation history
is of importance. The different behaviour of deformable materials in this simple shear can
be used to classify these materials.
If at a constant shear stress, t, the (shear) strain, y, becomes instantaneously constant, the
deformable material is called elastic rigid body or Hookean solid. If the (shear) strain, y,
does not become instantaneously but asymptotically constant, the material is called
viscoelastic solid.
The linear relationship

't = Gy (3.3)
is called Hooke's Law for solids. The constant ,G, is called rigidity modulus.
A liquid is characterized by the fact that at a given shear stress the (shear) strain does not
become constant. Instead of this, the strain rate, which in a pure shear flow is a shear rate,

. x dx d(tga)
y=-=-=--
h dy dt (3.4)

becomes constant. In this case of a pure shear flow material surfaces slide relatively to
another in the direction of the force and without a stretching deformation. If eq. (3.4) also
y,
holds for the instantaneous values oft and the fluid is a pure viscous fluid. If for a given
shear stress, t, the shear rate asymptotically approaches a fixed value, the material is called
viscoelastic fluid.
The linear relationship

t = ni' (3.5)

is called Newton's Law for liquids and the material a Newtonian fluid. The constant, 11. is
called dynamic shear viscosity. If confusion can be excluded, it is often simply called
viscosity. The SI-unit of viscosity is the Pascal-second, abbreviated as Pas.
There are two limiting cases of material behaviour: the linear models of Hooke's Law
(1678) and Newton's Law (1687). Hooke's Law is the starting point for the small-strain
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 35 .

elasticity theory for elastic rigid bodies, whereas Newton's Law is the basics for the
Navier-Stokes-equation (see Chapter 4.1) for fluid mechanics. They are outside the scope
of rheology. Rheology deals with materials exhibiting elastic and viscous material
properties. Thus rheology is situated between the two classical cases, the elasticity theory
and the Newtonian fluid mechanics.According to Bingham (1929) "rheology is the study of
the deformation and flow of matter". This definition widely covers different materials, such
as plastics, rubber, asphalt, suspensions, lubricants, polymer and surfactant solutions etc.
The flow behaviour of gases and low molecular weight liquids, e.g. water and organic
solvents, can be described by Newton's Law, eq. (3.5). Fluids containing large molecules,
e.g. polymer solutions or melts, supermolecular structures, e.g. micelles, or supensions of
solid particles, exhibit a more complex flow behaviour, which is generally called flow
behaviour of non-Newtonian fluids, or when abbreviated, non-Newtonian behaviour.
Since the flow behaviour of the non-Newtonian materials mentioned above can be quite
different, the behaviour in a simple shear flow can be used to classify these materials
according to the behaviour of their shear-dependent viscosity.
If the shear viscosity decreases with increasing shear rate, the fluid is called shear-thinning
or pseudoplastic. Typical materials exhibiting shear-thinning, as we will see later on, are
polymer solutions.
If the shear viscosity increases with increasing shear rate, the fluid is called shear-
thickening or dilatant. Typical materials exhibiting shear-thickening in a certain shear range
are surfactant solutions and suspensions at high solid concentrations. These materials often
also exhibit shear-thinning at low shear rates.
So far materials have been considered which exhibit a time-independent shear viscosity at a
given shear rate. However, there are materials whose shear viscosity increases or decreases
with the time of shearing at a given shear rate, i.e. their flow behaviour is time-dependent.
If the shear viscosity at a given shear rate decreases with time, the material is called
thixotropic. This behaviour can be caused by the destruction of an internal structure, due to
shear.
If the shear viscosity at a given shear rate increases with time, the material is called anti-
thixotropic or rheopectic. This behaviour can be caused by the formation of internal
structures due to shear.
The measuring techniques for the shear viscosity can be found in Chap. 3.2.
Viscoelastic fluids can exhibit, as their name already indicates, viscous and elastic material
properties. Thus a given material can behave more viscous or more elastic, depending on
the time scale of the deformation. At small strain rates the viscous material properties
dominate, whereas at high strain rates the elastic material properties dominate, regardless of
whether the strain is caused by a shear or an elongational deformation. In the intermediate
36 CHAPTER III

range both material properties are of importance. Thus for characterizing the flow behaviour
of viscoelastic fluids a dimensionless number, the Deborah number,

9
De=-
t (3.6)

is introduced with e as a characteristic time of the material, the so-called relaxation time,
and t a characteristic time of the deformation, due to the flow.
In order to achieve information on the influence of the time scale on viscoelastic material
properties it is convenient to subject the material to an oscillatory shear. As a response the
shear stress will also oscillate at the same frequency, but a phase shift between the shear
and the shear stress will occur. This phase shift is 0= oo for a pure elastic fluid and 8= 900
for a pure viscous fluid. The phase shift of a viscoelastic fluid lies between these two
extreme cases. Mathematically, this reads for a sinusoidal oscillation:

r ='Yo sin( rot) (3.7)

and

't = 'to sin( rot+ li) (3.8)

If the stress response is decomposed into two waves of the same frequency, one in phase
(8 = 0) with the strain and one out of phase (8 = 90) with the strain, one achieves

't = 't' + 't" ='to sin(rot) +'to cos( rot) (3.9)

and

't"
tgo = ..J!.
'to (3.10)

In analogy to Hooke's Law, eq. (3.3), frequency dependent dynamic moduli can be defined
as

G' ='to
'Yo (3.11)

and
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 37

G" = 'to
'Yo (3.12)

With these definitions eq. (3.10) becomes

1:" G"
tg~=_Q_=-
'to G' (3.13)

G' is called storage modulus. It is a measure of the reversible elastic deformation energy
which is stored in the material. G" is called loss modulus or dynamic rigidity and is a
measure of the irreversible dissipated energy due to the flow of the material.
In analogy to Newton's Law, eq. (3.5), a complex viscosity, Tl*, can be defined as the
ratio of the shear stress to the shear rate

1:( t) = t( y( t) (3.14)

This complex viscosity, 11*, can also be decomposed into

1:" G"
t]'=~
'Yo m (3.15)

and

, 'to
11=-=-
G'
Yo m (3.16)

11' is called dynamic viscosity. 11" gives the elastic part of the complex viscosity, Tl*
An analysis of oscillatory data with equations (3.7) to (3.16) is only correct if the
deformation lies in the linear-viscoelastic region where the shear stress varies linearly with
the (shear) strain, y. Therefore the amplitude of the (shear) strain has to be small. Modem
rheometers are available in which the output and input signals for oscillatory shear
experiments are directly processed to achieve the dynamic material properties mentioned
above.
When a Newtonian fluid leaves a capillary of diameter, d, into air, no dramatic change in
the diameter of the jet occurs. In contrast, in a flow of a viscoelastic fluid the diameter of the
jet can inc~:ease up to three times d. This viscoelastic flow phenomenon is called die swell

effect. The physical reason for such a behaviour is that in the capillary flow of the
38 CHAPTER III

viscoelastic fluid an extra stress along the streamlines exists. When the fluid leaves the
capillary, these normal stresses in flow direction relax, the fluid contracts in axial direction
and expands in radial direction. This means that in a capillary flow, which is a shear flow,
viscoelastic fluids exhibit normal stresses, in contrast to Newtonian fluids.
Usually, see Fig. 3.2, the components O"nx, O"ny, and O"nz are given to characterize the
stress distribution in a flow. In this notation the first index provides the orientation of the
area and the second the direction of the force.
The components O"xx, cryy. and O"zz are called the normal stresses, whereas the components
O"xy, O"xz, O"yx etc. are called shear stresses. Due to symmetry reasons O"ij=O"ji Fig. 3.2
shows the stress components at a cubic volume element.

(J~ CJzy
CJzx
z

-~
~t(Jxx CJ+ ~

~ - CJyz
CJxz ~ _

~
~
CJyx

Figure 3.2 The components of stress at a cubic volume element

In a simple shear flow (see Fig. 3.1) of a viscoelastic fluid the following stresses occur:

't = Gxy = TJ(l)l (3.17)

(3.18)

and

(3.19)
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 39

N 1and Nz are called the first and the second normal stress difference. In a simple shear
flow of Newtonian fluids N 1 and Nz are zero. The occurrence of the normal stress
differences in simple shear flows of viscoelastic fluids is due to non-linear effects. The
physical origin for the occurrence of normal stress differences lies in the fact that the
internal microstructure of the fluid becomes anisotropic. For example, when a certain shear
rate is exceeded, a polymer molecule, which has a spherical shape at rest, becomes
deformed to an ellipsoid. These ellipsoids become aligned in flow direction with increasing
shear rate. Modem, commercially available rheometers are able to measure N1 and
sometimes also Nz.

Figure 3.3 Uniaxial elongational flow

So far we have considered the behaviour of viscoelastic fluids in simple shear flows.
Besides of this, also shearfree flows exist, so-called extensional or elongational flows. If
we consider the deformation of a cube, as shown in Fig. 3.3, we can formulate in analogy
to eq. (3.5), a relationship between the normal stress difference, cr, and the strain rate,
which is called elongational rate , E, in elongational flows

(J = CJxx -CJYY = CJxx -CSzz = -()-()au


Tl E E = Tl E dx
(3.20)

Here T1 is the elongational viscosity. Since the flow is symmetric around the x-axis and the
equation of continuity holds

()v dw l()u 1.
-=-=---=--E
()y ()z 2 ()x 2 (3.21)
40 CHAPTER III

The distinguishing feature of steady shearfree flows in comparison to shear flow is that
neighbouring particles move relatively to another at an exponential rate, whereas in steady
shear flows the separation of fluid elements takes place at a linear rate. Thus it can be
expected that fluids exhibit a totally different material behaviour in shear and elongational
flows where internal structure becomes anisotropic due to the deformation, see Bird et al.
(1987).
For a Newtonian fluid it follows:

(3.22)

This result shows that for a Newtonian fluid the extensional viscosity in uniaxial extension
is three times the shear viscosity, Trouton (1906). TR is the Trouton ratio. For a
viscoelastic fluid this ratio can increase considerably and in general is a function of the
elongational strain rate. For polymer solutions Trouton ratios up to 1o4 have been reported.
In analogy to the definitions of rheological terms in simple shear flows a fluid is called
tension-thickening ifii increases with increasing E, or tension-thinning ifl; decreases with
increasing . Polymer solutions of flexible molecules are often tension-thickening, whereas
polymer solutions of relatively stiff, rod-like polymers are tension-thinning.
This and the following chapter do not purport to offer a course in rheology and rheometry.
The interest is restricted to those subparts of rheology and rheometry which are or may be
important for describing the rheological behaviour of drag reducing fluids. In order to
achieve this, all we needed was a survey of some relevant material properties. The reader
who would like to go into more detail should look through the various textbooks on
rheology and rheometry, e.g. Barnes et al. (1989), Bird et al. (1987), or Tanner (1985).

3.2 Rheometry

3.2.1 Shear viscometers

Nowadays it has become easy to measure the shear viscosity behaviour of an unknown
material, because many commercially available viscometers exist for this p~rpose. The most
common are rotational viscometers. Rotational viscometers are produced with different
geometries. The simplest geometry, see Fig. 3.4, is the flow of a test liquid in the gap
between two concentric cylinders produced by rotating the inner or the outer cylinder. This
type of flow is called Couette flow.
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 41

torque
transducer

Figure 3.4 Couette viscometer

If the gap is narrow and the geometry of the system is given by the two radii, R1 and R2,
and the height, H, the shear rate, y, is given for a stationary flow by

(3.23)

with Q being the angular velocity of the rotating cylinder. When the torque M of the
cylinder is measured, the shear stress results in

(3.24)

and the shear viscosity in

(3.25)
42 CHAPTER III

In this approximation end effects at the liquid surface and at the bottom of the instrument,
are neglected. There are two ways of producing the flow: either the inner or the outer
cylinder is driven. Instruments based on the first method are called Searle-viscometers and
the ones based on the second method Couette-viscometers. The disadvantage of Searle-
viscometers is that due to centrifugal forces a secondary flow with so-called Taylor vortices
can be introduced at high shear rates. Using rotational viscometers the shear viscosity
behaviour of a material can be measured in the shear rate range from 10-6 to 1Q3s-1.

torque
transducer

_,
~t.-

Figure 3.5 Cone-and-plate viscometer

Another geometry of a rotational viscometer is the cone-and plate-geometry in which the test
fluid is filled between a rotating plate and a fixed cone of radius R, see Fig. 3.5. For this
geometry the shear rate of a stationary flow in the gap is constant

. Q
y=-
p (3.26)

with n being the angular velocity of the rotating plate, provided the angle, J3, between the
cone and plate is small. By measuring the torque, M, the shear stress
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 43

3M
't=--
2xR3 (3.27)

and the shear viscosity

(3.28)

can be obtained.
At high rotational rates secondary flows can influence the measurements. In theory, the tip
of the cone should contact the centre of the plate. However, in practice the cone is often
truncated.
Very similar to the cone-and plate-viscometer is the geometry of the parallel-plate
viscometer, see Fig. 3.6. The shear rate of the torsional flow of the test liquid in the gap
with a height, h, between the two parallel plates with a radius, R, is

. rn
y=-:..&
h (3.29)

torque
transducer

Figure 3.6 Parallel plate viscometer


44 CHAPTER III

Here, the shear rate becomes non-homogeneous and increases linearly to the radius of the
plate. It beomes a maximum at the rim of the plate, r = R. In general, when the torque, M,
is measured, the shear stress is given by

1:(" )==~[
3
3 + d(lnM)]
'YR 2xR d(lnyR)
(3.30)

with

. R
'YR ::::-Q
h (3.31)

being the shear rate at the rim of the plate. Consequently, the shear viscosity becomes

(. ) Mh
11 'YR :::: ZxR4Q[ + d(lnM) ]
3
d(lnQ) (3.32)

The parallel-plate geometry and the cone-and plate-geometry can often be mounted
alternately in commercial viscometers. Sometimes both geometries can also be installed in
Couette-viscometers.
Since the maximum shear rate achievable in Couette-viscometers is limited due to the
occurrence of Taylor vortices or the transition to turbulence, for higher shear rates only
capillary viscometers are available, in which the shear viscosity can be measured by using
the Poiseuille equation. It is important to note that for measuring shear viscosities of non-
Newtonian fluids by capillary viscometers it has to be considered that the shear rate at the
wall

== 4Q [~+ 1 d{lnQ)]
'Yw xR 3 4 4 d(ln'tw)
(3.33)

is modified. In this equation Q stands for the flow-rate, R the radius of the capillary, and 'tw
the wall shear stress. The term in brackets in eq. (3.33) is called Rabinowitsch correction.
The first normal stress difference, N 1, can be obtained in a cone-and plate-viscometer (see
Fig. 3.5) by measuring the normal force, F, on the plate

2F
N1 ::::--2
xR (3.34)
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 45

Measuring the force, F, on one plate in the parallel-plate viscometer (see Fig. 3.6) provides
the difference between the ftrst and second normal stress

N -N =_!_[I+ 1 d(InF) ]
I 2 xR 2 2 d(lnyR) (3.35)

Thus force measurements in a cone-and plate- and in a parallel-plate-viscometer can be


combined in order to achieve N 1 and Nz separately. However, since Nz is normally smaller
than N 1, usually one order of magnitude less, these measurements have to be very precise.

3.2.2 Elongational Viscometers

The measurement of elongational viscosities is one of the most difficult problems in


rheology, especially for low-viscosity fluids. Several extensional viscometers have been
proposed. However, in most situations no steady state can be reached in these instruments
and only transient elongational viscosities, depending on the residence time in the
elongational fteld, can be measured. Nevertheless, in many flow situations, as in the local
elongational ftelds of a turbulent flow, no steady state is reached either. Furthermore, since
the actual flow behaviour of viscoelastic materials depends on the deformation history, the
flow in an elongational rheometer should be shearfree and the strain history of the test
material in the rheometer should have no influence. Nevertheless, the elongational ftelds in
turbulent boundary layers are not shearfree either. In many practical flow situations also
combined shear and elongational ftelds exist too.
The difficulties in measuring elongational viscosities of polymer solutions and the state of
the art of the measuring techniques are described in a special issue of the Journal of Non-
Newtonian Fluid Mechanics, Vol. 35 (1990) 85-470. Several methods which are suitable
for polymer melts, as the homogeneous stretching method, Meissner (1972), or the
constant stress device, Mtinstedt (1975), cannot normally not be used for polymer and
surfactant solutions exhibiting only low shear viscosities.
In the following the main experimental methods for determiningg elongational properties of
polymer and surfactant solutions will be briefly reviewed. The most common techniques are
spinning techniques. In the tubeless- or open-syphon-technique the test fluid which is at
rest in an open reservoir is drawn up by a vacuum pump, Astarita & Nicodemo (1970), or
by a wind-up device, like a driven rotating cylinder, Martischius (1982). The force for
drawing up the fluid and the flow-rate are measured. When the contour of the drawn
polymer thread is recorded by a (video) camera, the elongational viscosity and the
46 CHAPTER III

elongational rate can be obtained. Nevertheless, although all boundaries of this flow, which
is also called Fano-flow, are free surfaces, i.e. no shear causing wall effects is involved,
the flow itself is not shearfree, as shown by Matthys (1988). The shearfree surfaces cause
shearing inside the drawn-up thread, which can be shown by a stress balance of a fluid
volume close to the surface of the reservoir.
Similar to the method of the tubeless syphon is the measuring method of the spin-line-
rheometer, Chang & Denn (1979). The fluid is drawn out of a reservoir through an orifice
by means of a suction or a wind-up device. The data acquisition and analysis is also similar.
Besides the shearing due to the shearfree surface the test fluid may still remember the
straining in the orifice, i.e. the deformation history can influence the measured elongational
properties of the test fluid. Spin-line-rheometers are commercially available.
A stagnation point flow which is shearfree is used in the elongational rheometer of Fuller et
al. (1987). Fluid is sucked into two opposed entry tubes and the force onto one of the tubes
is measured. The only disadvantage of the instrument is that the residence time of the fluid
particles in the elongational field is not constant. When the instrument is operated in the
opposite direction, i.e. the test fluid is pushed out of the tubes and two opposing jets are
produced, the influence of a pre-shearing on the elongational properties of the test fluid can
also be studied. This instrument is commercially available.
The easiest way to obtain information on the elongational properties of a test material is to
study its behaviour in contraction flows. By measuring the flow-rate versus the pressure
drop in a contraction one can achieve an estimate of the elongational viscosity as a function
of the elongational rate. However, the interpretation of the data may become difficult when
entrance vortices develop upstream from the contraction, causing an additional pressure
drop and changing the kinematics of the flow field. The flow field in a contraction is
inhomogeneous. On the axis the flow is almost purely elongational, whereas a shear flow is
realized near the wall. Between these two extreme situations mixed flows exist in which the
ratio from shear to elongational stresses changes continuously. The magnitude and duration
of the elongation are therefore not the same for all fluid elements entering the orifice.

3.3 Rheology of dilute polymer solutions

In a dilute polymer solution the polymer molecules exist as isolated macrom~lecules which
become solvated chaotic coils. With the quality of the solvent the size of the coils increases.
In a dilute solution the dissolved molecules do not touch one another and there exist no
forces between the molecules. Hence, there are no intermolecular interactions. The critical
concentration c *, the limit between the dilute and the concentrated solution, can be
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 47

calculated from light scattering measurements. Besides - as we will see later on - also
viscometric measurements can be used.
When polymers are dissolved in a solvent, the viscosity of the solution increases with the
polymer concentration. When the polymer concentration is kept fixed the solution viscosity
increases with the molecular weight, M, of the polymer. Thus it makes sense to define a
relative viscosity as

_ llo
llrei--
lls (3.36)

which is the ratio of the viscosities of the polymer solution and the solvent. In an idealized
form the polymer molecules in solution are coils of spherical shape. Thus, for dilute
solutions the equation of Einstein (1906), eq. (3.37), which was originally developed for
suspensions of spherical particles, can be used to describe the viscosity increase:

110 =11s(1+2.51) (3.37)

where <I stands for the volume fraction of the spherical particles. For a polymer solution <I
would be the .ratio of the volume of the dissolved molecules to the volume of the solution. If
the polymer concentration of the polymer solution, c, is known, eq. (3.37) can be rewritten
as

(3.38)

where Po is the density of the polymer without the solvent. In order to elucidate the polymer
contribution to the viscosity often a reduced viscosity

,... _ llrel-1
red-
c (3.39)

is used. With eqs. (3.36 - 3.38) eq. (3.39) becomes

2.5
llred = -
Po (3.40)

If no intermolecular interactions exist, i.e. the solution is dilute, (c~O), and the
deformation has no influence on the viscosity, (y~O), i.e. the deformation is in the range
48 CHAPTER III

of zero-shear viscosity, 11red. is called intrinsic viscosity, [11]. The intrinsic viscosity, [11],
increases with the molecular weight. In the dilute regime a linear relationship between the
reduced viscosity, 11red. and the polymer concentration, c, exists. When the zero-shear
viscosity, i.e. the shear viscosity at extremely low shear rates, (y~O), is measured in this
dilute regime this linear relationship can be used to calculate the intrinsic viscosity by
extrapolation, (c~O).

Between the molecular weight, M, of a polymer and the intrinsic viscosity, [11], an
empirical relationship exists

(3.41)

In this equation the constant, K, depends on the polymer-solvent system. The exponent, a,
is called Mark-Houwink exponent and is found to vary between 0.5 and 2. If the constant,
K, and the exponent, a, of a particular polymer-solvent system are known, this relationship
can be used to determine the molecular weight of a polymer sample by viscosity
measurements. These constants are tabulated for various polymer-solvent systems, e.g. by
Brandrup & Immergut (1975) or for water soluble polymers by Bekturov & Bakauova
(1986). The constants, Kanda, depend on the molecular weight distribution of the polymer
and on temperature. The exponent, a, provides information on the shape of the molecules in
solution and on the polymer-solvent interactions. In an ideal solvent an equilibrium between
the solvation forces of the solvent and the agglomeration forces of the chain segments of the
polymer exists which results in a force-free polymer coil. An ideal solvent of this kind is
called a fJ-so/vent. Under 8-conditions the constant, a, is found to be 0.5 for flexible linear
polymers. The constant, a, of flexible polymers in non-S-solvents lies in the range of 0.5 <
a< 0.8. Polymers exhibiting values of a> 0.8 are usually suspected to be semi-flexible and
a solution of rigid rods would exhibit a value of a = 2. The method of determining the
molecular weight via the intrinsic viscosity cannot be used for solutions of polyelectrolytes
in distilled water, because due to the charge of the polymer chains the polymer coil expands
when the solution is diluted. This holds for anionic or cationic polyacrylamides, which are
frequently used in drag reduction experiments.
Most polymer solutions exhibit shear-thinning (see Chapter 3.1) in a steady simple shear
flow. In Figure 3.7 a typical example is given. The polymer, WSR 301, a
polyethyleneoxide of Union Carbide, according to the manufacturer has a molecular weight
ofM = 4 106 g/mol. It is frequently used in drag reduction studies. For all concentrations
the shear viscosity becomes constant at low shear rates. This constant viscosity at low shear
rates is called zero-shear viscosity. At high polymer concentrations this region of zero-shear
viscosity, sometimes also called lower Newtonian region, is followed by a shear-thinning
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 49

region at higher shear rates. The shear-thinning becomes more pronounced with increasing
polymer concentration. In the experimental results shown in Fig. 3.7 the upper limit at high
shear rates of this shear-thinning region could not be detected. Nevertheless, all molecular,
phenomenological, or theoretical models predict the existence of a second, upper
Newtonian region at high shear rates in which the shear viscosity once more becomes
constant. Due to experimental problems, such as the degradation of polymer molecules at
high shear rates, temperature effects due to dissipation and deviations from a parallel flow
due to elastic effects, it is difficult to obtain reliable experimental data in this high shear rate
region for solutions of flexible polymers, like polyacrylamide or polyethyleneoxide.
Sometimes also a shear-thickening region is detected at high shear rates, e.g. Wolff (1982)
or Vrahopoulou & McHugh (1987), which follows the shear-thinning region at lower shear
rates.
There are many empirical and constitutive equations which predict the shape of the shear
viscosity behaviour of polymer solutions. Since the shear-thinning becomes more
pronounced at higher polymer concentrations, in the semi-dilute or in the concentrated
range, they will be described in the following chapter.

..............................................


...
..........
<> <>
0 0<><> <><> <><> () <> <> <> <><> <>:
<>$$0 0
0
<>oooo
ooooooooo~oooooooooooo
ocoooo
6AAAAAAAA.o.AAAAAAA.o.AAA AA4A.o..o.
xxxxxxxx ~~x~xxxxxxx
~~~~~$i~ ~.~i~~~~ ~~~~~~

10 1

Figure 3.7 Shear viscosity of WSR 301-solutions of different concentrations, from


Vissmann & Bewersdorff ( 1990)

At low polymer concentrations (see Fig. 3.7) the shear-thinning behaviour exhibited by
almost all polymer solutions, can be neglected. Fig. 3.8 elucidates this showing the shear
viscosity behaviour of a polyethyleneoxide solution, WSRN12K, from Union Carbide,
50 CHAPTER III

with, according to the manufacturer, an average molecular weight of M = 1o6 g/mol.


Solutions of this polyethyleneoxide exhibit drag reduction in the turbulent flow regime,
even at the lowest concentration of 20 ppm studied. Up to a concentration of 2000 ppm no
shear thinning behaviour can be observed. Thus the shear-thinning cannot be the
rheological property which is responsible for drag reduction in the turbulent flow regime.
By comparing the shear viscosity behaviour of the two polyethyleneoxides in Figs 3.7 and
3.8 the strong influence of the molecular weight becomes evident.
The elastic properties of viscoelastic fluids manifest themselves in a non-vanishing first
normal stress difference in a steady shear flow. The conventional rheometers for measuring
the first normal stress difference, usually by a cone-and-plate rheometer (see Chapter 3.2.1,
Fig. 3.5), are not suitable for dilute polymer solutions, because in low-viscosity fluids the
fluid flows out of the gap, due to centrifugal forces, even at low shear rates. Furthermore,
the normal stresses are too small to be resolved by the force transducers of commercial
rheometers. Thus available results on the behaviour of the normal stress differences in shear
flows are restricted to those polymer solutions whose solvent viscosities are high or whose
polymer concentrations are higher than those needed to produce drag reduction in the
turbulent flow regime. These measurements will be reviewed in the following chapter.
In Chap. 3.1 it was shown how oscillatory shear flows of low amplitude can be used to
determine the elastic properties of a sample. In a simple shear flow the deviations from
Newtonian material behaviour manifest themselves in the viscometric functions, 11. N I and
N2, whereas in an oscillatory shear flow deviations in the viscometric functions, 11' and G',
occur. Thus it is not surprising that relationships between these material functions exist. At
low frequencies and shear rates

(3.42)

and

G'(m)l = NI(Y)I
2 2'Y.2
(i) (l}---)0 y---)0
(3.43)

Due to eq. (3.42) the dynamic viscosity, 11', for ro~O. becomes identical with the zero-
shear viscosity 110 for y~O. Eq. (3.43) provides a relation between the storage modulus,
G', and the first normal stress difference for low frequencies and shear rates, respectively.
In many cases, especially for low viscosity fluids, it is easier to do oscillatory experiments
instead of realizing a steady shear flow at extremely low shear rates. In these cases 11 and
N I are calculated from 11' and G'. The shear viscosity, 11, as well as the dynamic viscosity,
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 51

TJ', normally decline with y or ro, respectively. It has been found that many but not all
polymer solutions and polymer melts follow the so-called Cox-Merz-rule, Cox & Merz
(1958). According to Cox & Merz the viscometric functions, TJ= f('y) and TJ*= f(ro), see eq.
(3.14), are identical. Due to the existence of such a relation oscillatory experiments are
frequently used in literature in order to characterize the elastic properties of dilute polymer
or surfactant solutions.

"~! '""' I I"" I I I '""I 1 illiilj I I I lllllj I l:::j


...

f-

~
0 ooooooooo~oooooooooooo

oooooo
oooooo
i:J.AAAAI::.

~~~~~~
*

10 10'

Figure 3.8 Shear viscosity of WSRN12K solutions of different concentrations, from


Vissmann & Bewersdorff ( 1990)

Recently, Tirtantmadja & Sridhar (1993) used this method for polymer solutions and
developed a filement stretching device for measuring the elongational viscosity of low-
viscosity fluids. The fluid sample is held between two discs which move apart at an
increasing velocity so that the elongational rate becomes constant. This was the first time
that a steady state value of the elongational viscosity could be evaluated for a concentrated
polymer solution. The steady state Trouton ratios varied between 2- 5103.
Common methods for determining the elongational behaviour of (dilute) polymer solutions
are the behaviour in orifice flows, the spin-line rheometer, the opposing jet technique (see
Chapter 3.2), and for a first approximation the behaviour in porous media flows. The first
and the last method have the disadvantage that the elongational fields are inhomogeneous.
Thus the elongational viscosity cannot be exactly determined as a function of the
elongational rate. The available experimental data for dilute polymer solutions show that the
Trouton ratio, eq. (3.22), can be considerably increased. The elongational viscosity
increases with the elongational rate, i.e. the solution behaves tension-thickening, until at
52 CHAPTER III

high elongational rates an asymptotic value is approached. This asymptotic value of the
Trouton ratio increases with the polymer concentration. Fuller et al. (1987) found an
asymptotic Trouton ratio of 11.9 for a 25 ppm polyacrylamide solution, which increased up
to 355 when the polymer concentration was raised up to 500 ppm. Solutions of the more
rigid polysaccharide xanthan gum exhibited remarkable lower Trouton ratios in the same
concentration range.
It is characteristic of viscoelastic fluids that the flow behaviour depends on the deformation
history. In many experimental techniques for determining the elongational behaviour of
polymer solutions (see discussion in Chapter 3.2.2) the influence of pre-shearing on the
measured elongational viscosity cannot be excluded. On the other hand, its influence was
demonstrated in many experimental studies. For example, Oliver & Ashton (1976)
examined the elongational behaviour of polymer solutions in a triple-jet-apparatus. They
varied the pre-shearing by mounting various cylindrical obstacles in the capillary. The
elongational viscosity of polyethyleneoxide solutions could be doubled by pre-shearing.

11

Figure 3.9 The modified rotational viscometer for studying the influence of pre-shearing on
the elongational behaviour, from Vissmann & Bewersdorff (1990)

A first attempt at quantitatively estimating the influence of pre-shearing on the elongational


behaviour of dilute polymer (and surfactant) solutions was made by James et al. (1987) and
Wunderlich & James (1987). The polymer solutions were first subjected to a planar,
laminar, channel flow, and a small portion of the main flow was then directed through a
converging channel. The wall shear rate was used to characterize the pre-shearing. This is
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 53

an approximation, because in channel flows the shear rate is a function of the distance from
the wall. Not only fluid from the near-wall region is diverted through the converging
channel. When the pre-shearing exceeded a certain value the elongational viscosity
estimated from the pressure drop in the converging channel increased and strain hardening
occurred.
This method was improved by Vissmann & Bewersdorff (1990). A Couette flow instead of
a channel flow was chosen because the shear gradient in Couette flows is constant. All
material in the entrance flow of the orifice therefore has the same pre-shearing deformation
history. Fig. 3.9 shows the experimental set-up. Measured parameters are the angular
velocity of the rotating cylinder, the flow rate through, and the pressure drop over the
orifice. The flow rate is determined from the increase in weight per time.
By analyzing the pressure drops of a Newtonian fluid and a dilute viscoelastic polymer
solution, which exhibits no deviations from Newtonian behaviour in simple shear and pure
elongational flows, the transient elongational viscosity of a dilute polymer solution can be
estimated. Fig. 3.10 provides an example of the elongational behaviour of a dilute
polyethyleneoxide solution. Here, the relative elongational viscosity, the ratio of the
elongational viscosity to three times the shear viscosity, which is one for a Newtonian fluid
in an uniaxial' elongational flow, is plotted versus the Deborah number, De, (see eq. 3.6)
which is the product of the Rouse relaxation time and the elongational rate of the orifice
flow. At low pre-shear rates, y, no deviations from Newtonian behaviour occur. Only
when a critical shear rate is exceeded, which is characteristic for each polymer-solvent
system and independent of the elongational rate, the elongational viscosity increases. The
dotted lines in Fig. 3.10 connect data points which were obtained at a constant pressure
drop across the orifice. With increasing Deborah number the relative elongational viscosity
first increases, then reaches a plateau value, and at the highest elongational rate it decreases
again. It seems that at high elongational rates the influence of pre-shearing declines. By
increasing the molecular weight of the polymer the critical pre-shear rate is lowered. Thus it
seems that the increased elongational viscosity is mainly caused by the deformation and
orientation of the polymer molecules in the shear flow. Calculations of the alignment angle
of the polymer molecules using the formula of Cottrell et al. (1969) showed that the
polymer molecules start to align at the critical shear rate. At the maximum value of the
elongational viscosity they are nearly completely aligned in flow direction of the shear flow.
Nevertheless, the above method has restrictions. Entrance vortices can occur at the orifice.
The presence of entrance vortices results in a higher apparent elongational viscosity. Thus it
is necessary to carry out a flow visualization of the entrance flow at the same shear and
elongational rates in order to check whether entrance vortices are present or not.
The fact that the elongational viscosity of dilute (drag reducing) polymer solutions depends
on the pre-shearing also helps to understand the behaviour of these solutions in turbulent
54 CHAPTER III

shear flows (see Chapter 5). Here they exhibit the so-called onset behaviour, which means
that below a certain critical wall shear stress their turbulent friction behaviour does not differ
from that of the solvent. Drag reduction in turbulent flows occurs when a critical wall shear
stress, or more accurately, a critical wall shear gradient is exceeded. In the near-wall region
of a turbulent flow the elongation occurs in a similar way as in the modified rotational
viscometer shown above. In both cases the polymer solution is already subjected to a shear
gradient before experiencing the additional elongational field.

10
symbol

8 . '' '
''
''
.. '' ''
''
I 8 a

~~I~
4

0
0 .. 10 15 20
De [-J

Figure 3.10 Relative elongational viscosity of a 15 ppm WSR 301-solution as a function of


the Deborah number for different pre-shear rates, from Vissmann &
Bewersdorff ( 1990)

So far we have seen many flow situations where even a dilute polymer solution may exhibit
drastic differences between its rheological behaviour and the Newtonian behaviour. On the
other hand, in a simple shear flow the shear viscosity behaviour of dilute polymer solutions
exhibts, if detectable, negligible deviations from Newtonian behaviour. The aim of one
branch of rheology is to develop a constitutive equation to describe such a flow behaviour.
A (rheological) constitutive equation is the relation between stress, strain and time and
sometimes other variables such as temperature. Generally in the average case the stresses in
the material at the actual time are related to the actual and previous deformation experience
of the material, i.e. the influence of the deformation history on the actual stress can be
considered.
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 55

3.3.1 Constitutive equations

The simplest constitutive equations are the ones of a Newtonian fluid, eq. (3.5), and a
Hookean solid, eq. (3.3). They can also be considered as mechanical models of the
material. The Newtonian flow is represented by a dashpot, an element in which the force is
proportional to the rate of extension, whereas the Hookean deformation is represented by a
spring, an element in which the force is proportional to the extension. Since viscoelasatic
fluids exhibit viscous and elastic material properties, there have been many attempts to
develop mechanical models for viscoelastic fluids by arranging dashpots and springs
parallel or in series.

Figure 3.11 a&b: (a) Mechanical model of a Maxwell fluid (b) Mechanical model of a
Kelvin solid

Well-known mechanical models obtained by such arrangements of dashpots and springs are
the Maxwell model (dashpot and spring in series, see Fig. 3.1la) and the Kelvin model
(dashpot and spring parallel, see Fig. 3.1lb).
For more complicated materials more than two of these elements can be combined, as was
done in the Jeffreys (see Fig. 3.12) and in the Burgers model (see Fig. 3.13). The
behaviour of these model fluids in rheological flows is described in the textbooks on
rheology (see Chapter 3.1).
However, it has been found that the rheological behaviour of dilute polymer solutions is
difficult to describe when using such continuum approaches in order to find a suitable
constitutive equation. It may happen that such a constitutive equation describes the
behaviour, in one flow situation well, but totally fails in another flow situation. If the
microstructure of the fluid is taken into account, i.e. if a physical model for the polymer
molecule is considered, one may expect to gain a deeper insight into the viscoelastic flow
behaviour because viscoelastic flow phenomena are mainly caused by intramolecular forces
56 CHAPTER III

of the polymer molecules. The reason for the occurrence of these intramolecular forces is
the orientation of the chemical-bond vectors in the polymer chain. This orientation changes
when the surrounding solvent fluid is' deformed.

Figure 3.12 Mechanical models equivalent of a Jeffrey fluid

Figure 3.13 Mechanical models equivalent of a Burgers fluid

In Chapter 2 we already considered the molecular models of Zimm and Rouse, which are
bead-spring models. The beads exhibit fluid friction whereas the springs are considered to
be frictionless. When a constitutive equation based on the Zimm-Rouse theory is developed
one achieves a linear viscoelastic model which is only suitable to conditions of small-
amplitude oscillatory shear flow.
Another type of models considers the polymer chain as a dumbbell, i.e. two beads are
located at the ends of the model molecule and connected by a spring. The frictional drag acts
on the two beads whereas the spring represents the elastic behaviour. Based on such
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 57

dumbbell models various constitutive equations for dilute polymer solutions can be
deduced. The spring can be considered as Hookean, with limited extension, or as finitely
extensible, non-linearly elastic. In the later case the so-called FENE dumbbell model is
obtained, which predicts shear-thinning and a non-vanishing first normal stress difference.
In general, it seems that these dumbbell models are the most realistic models for constitutive
equations of dilute polymer solutions. However, due to their complex mathematics their
practical use is sometimes restricted and only numerical solutions can be obtained for
practical flow situations. A review of these dumbbell models is given in the textbook of
Bird et al. (1987), see Chapter 3.1.
At present it is difficult, on the basis of rheological measurements, to decide which one of
these constitutive equations is the best for describing the flow behaviour of dilute polymer
solutions. As we have seen, the most dramatic deviations from Newtonian behaviour occur
in elongational flows. Thus, provided there were no difficulties in measuring steady-state
elongational viscosities, this type of flow would be an ideal candidate for testing the various
constitutive equations.
Due to the recent improvements in the measuring technique for determing the elongational
viscosity, Tirtaantmadja & Sridhar (1993), it can be expected to find an appropriate
constitutive equation for each polymer-solvent system in the near future.

3.4 Rheology of semi-dilute polymer solutions

As the polymer concentration increases, the probability of collisions between polymer coils
becomes more frequent and finally causes an overlapping and entanglement of the polymer
chains. The critical concentration at which the polymer molecules start to touch each other is
called overlap concentration, c*. When the polymer coil is considered to be a rigid, hard
sphere, the overlap concentration can be calculated if the radius of gyration, Ro. of a
polymer coil is known. If c < c*, the solution is called dilute, and vice-versa, if c > c*, the
solution is called concentrated. The overlap concentration of a polymer-solvent system
decreases with the molecular weight. In drag reducing polymer solutions, which usually
possess a molecular weight of M = 1()6 - 2107 g/mol, this concentration exists can be of
the order of 10- 100 ppm. Thus in these solutions a wide range of concentrations in which
c > c*. Nevertheless, their absolute concentration is quite low. According to the definition
mentioned above these solutions would be called concentrated although their absolute
concentration is small. Thus they are often referred to as semi-dilute.
Semi-dilute polymer solutions can exhibit a pronounced shear thinning behaviour in shear
flows. Fig. 3.7 provides such an example. At low shear stresses (or shear rates) the shear
viscosity becomes constant and is called zero-shear-viscosity, Tlo, see Chapter 3.3. The
58 CHAP1ERIII

regime of zero-shear viscosity is followed by a shear stress (rate) region in which the shear
viscosity decreases with the shear stress (rate). Usually a logarithmic scaling is used for
both axes. In such a plot a linear decrease is often found for polymer solutions. Hence, this
"power-law" region can be described as

n-1
1}= k y (3.44)

1) Rhodopol 23
[Po sl
Do
. 500ppm
co
u"oo
1000ppm

.... . .. '0,.
0 2000ppm

.. .
1

..
o~a

.......... .. .
c:iJ,
.. ... . 'rna

. ......... ..,.,.
2

COn
De

-
c

- -
10 2
T [Pol

Figure 3.14 Shear viscosity behaviour of various concentrated solutions of Rhodopol23, a


Xanthan gum from Rhone-Poulenc, from Bewersdorff & Berman (1988)

as originally proposed by de Waele (1923) and Ostwald (1925). Fluids exhibiting such a
behaviour are called power-law-fluids or Ostwald-de-Waele fluids. In this equation k is
called the consistency and n the power-law index. For polymer solutions n can vary
between zero and one. At high shear stresses the data of the Xanthan-gum solutions shown
in Fig. 3.14 tend to approach another asymptotic value of the shear viscosity, 'Tloo , the
viscosity at infinite shear stress (rate). The shape of the curve of the shear viscosity can be
descibed in the whole shear rate range by the Cross model (1965)

(3.45)
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 59

K and m are constants of the polymer solution. Besides the power-law and the Cross-model
there exists a variety of basically empirical models which can be used to fit the shear
viscosity behaviour.
The elastic properties of semi-dilute or concentrated polymer solutions manifest themselves
in a non-vanishing first normal stress difference. Fig. 3.15 provides a typical example for
the behaviour of the first normal stress difference, N 1. and the shear stress, 't, as a function
of the shear rate, y. Besides the region at low shear rates, the zero-shear viscosity regime,
the first normal stress difference is higher than the shear stress, a behaviour typical for
solutions of flexible polymers, with a slope of the first normal stress difference as a
function of the shear rate between 1 and 2.

0
0 '"" ""' ""
N1 , 1 g "~

[P~
0 "
0
0
" ll't
0
"
""
0
0 "
" "oDD
0
0
Do
oo
"
0
0 "" 0
0

0
0

8
0

__l_LW
" IIIII Ill!
0
0.01 0.10 1.00 1 o.oo 100.00 1 000.00

y [1/s]

Figure 3.15 Behaviour of the flrst normal stress difference and the shear stress as a function
of the shear rate for a 2% polyisobutene solution (Oppanol B200, BASF AG) in
decalin

Nowadays it is also possible to measure the behaviour of the second normal stress
difference, Nz, in shear flows. In general, it is found that the second normal stress
difference is negative or zero, which is predicted by the Weissenberg hypothesis. The ratio
of Nz to N 1 usually lies at about 0.1 or less.
60 CHAPTER III

3.5 Rheology of surfactant solutions

First, let us repeat some important physico-chemical properties of surfactant solutions from
Chapter 2.3. Due to the chemical stucture of the surfactants, i.e. their molecules consist of a
hydrophilic and a hydrophobic part, in water the surfactant molecules tend to form
aggregates, called micelles, when their concentration exceeds a critical value, the critical
micelle concentration, CMC. In a surfactant solution, which is in thermodynamic
equilibrium and whose concentration is above the CMC, there is always an equilibrium
between the micelles and single surfactant molecules. Micelles can contain up to 1,000
single surfactant molecules. In this way their size and form become compatible with that of
drag reducing polymer molecules. Depending on the dimensions of the hydrophilic and the
hydrophobic part of their molecules the shape of the micelles can be globular, disc-like or
rod-like. Since it has been found that drag reducing surfactant solutions always contain rod-
like micelles, the following remarks on the rheology of surfactant solutions will be
restricted to the rheology of (dilute) surfactant solutions containing rod-like micelles.
The diameter of a rod-like micelle is about twice the length of the extended hydrocarbon
chain of the surfactant molecules, typically 2 - 4 nm, whereas its length can vary between
25 and 250 nm in the concentration range used for drag reduction, usually 100- 2,000
ppm.
Thus one would expect a rheological behaviour of the surfactant solutions, similar to that of
fibre suspensions, which will be treated in the following Chapter, i.e. due to the anisotropic
geometry of the micelles orientation effects of the rods can occur in flow situations: Rigid
particles with an asymmetric shape, such as rods, can be oriented in shear flows. The
deformation rate duj/axj. can always be splitted into a symmetric and skew-symmetric part,
the strain rate dij. and rotation tensor rij- Due to this decomposition it is obvious that during
the rotation of the rod the velocity gradient at its end varies. Consequently, the rotational
speed of the rigid particle becomes non-uniform, an effect which becomes more
pronounced as the deformation (shear) rate increases.
At low shear rates the micelles rotate in a shear flow with a nearly constant rotational speed,
and there is only a weak alignment angle of 450 to the flow direction preferred. Since the
surfactant concentration normally lies in the dilute regime, in a shear flow the micelles can
rotate without hindering one another. When the shear rate increases, the rotational speed of
the rod-like micelles becomes non-uniform and the rods become more and more aligned in
flow direction. It is the shear which causes this alignment. In the opposite direction the
Brownian motion acts. Often the ratio of the shear rate, y, to the rotational diffusion
constant, Dr. which characterizes the Brownian motion, is used to describe the alignment
process of the micelles in the shear flow. At low shear rates the influence of the Brownian
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 61

motion dominates, whereas at high shear rates the influence of the shear gradient becomes
more pronounced and the rods are preferably oriented in flow direction. This means that the
rods still rotat in the shear flow, however, their most probable orientation is the one in flow
direction. The alignment is a dynamical process and therefore time-related to the flow field.
Due to this alignment process of the micelles the behaviour of the shear viscosity as a
function of the shear rate becomes very similar to that of semi-dilute or concentrated
polymer solutions. At low shear rates there is a region of zero-shear viscosity followed by a
shear-thinning region at higher shear rates. This shear thinning is due to the fact that aligned
rods cause a lower hydrodynamic resistance. If the micelles are considered to be rigid rods
and the surfactant concentration is in the dilute regime, i.e. no interaction between the
micelles exists, according to Doi and Edwards measurements of the zero-shear viscosity can
be used to calculate the length of the micelles

- 600 ppm
- 900 ppm
-1200 ppm
10- 2
;;

.. .
0
,g.
""


:: ...
10- 3
tttttt!t :::::tt.
10- 1 10 10 1 10 2 103 10 4
1 ls- 11

Figure 3.16 Shear viscosity behaviour of solutions of equimolar mixtures of


hexadecyltrimethylammoniumbromide and sodiumsalicylate in tap water at
T=200 C, from Vissmann & Bewersdorff (1990)

(3.46)

T)CMC is the shear viscosity at the critical micelle concentration where the formation of the
micelles starts, n stands for the number of rods per volume and L for the length of the rods.
In this equation, which is similar to Einstein's equation for the viscosity of a dilute
suspension of spheres, TICMC is entered in order to eliminate the influence of the viscosity
62 CHAPTER III

due to the dissolved single polymer molecules. Eq. (3.46) holds only in the dilute
concentration regime in which the rotational volumes of the rods do not overlap. If it is used
for calculating the length of the rods, accurate viscosity measurements are necessary
because the contribution of the rods to the zero-shear viscosity is relatively small.

0
0
I
~
start-up l--- relaxation

I
.I
I
I
..
0
0
~
0
.:.
u

\~ .
g
I n-
0o':--.-:-:oo::-o-----"2:-c.e=oo=------:-5-::.eo=o----':-
.-::o=o---'----',-,-._-=.o=---"""-c__,n.:.ot.oo

Figure 3.17 Start-up and relaxation behaviour of an equimolar mixture of


hexadecyltrimethylammomiumbromide and sodiumsalicylate in distilled water at
c = 1.8 mmol/1, T = 20 Oc andy= 5 s-1, to be published by de Buhr &
Bewersdorff

Fig. 3.16 provides the behaviour of the shear viscosity of a viscoelastic drag reducing
surfactant solution at three different concentrations. This solution contains rod-like micelles.
At low shear rates the shear viscosity seems to be constant and no shear-thinning behaviour
can be observed, i.e. the shear viscosity behaves Newtonian in this regime. The shear
viscosity of the solutions is only slightly increased, due to the dissolved surfactants.
However, when a critical shear rate is exceeded, the shear viscosity suddenly increases.
The critical shear rate as well as the absolute value of the shear viscosity do not only depend
on temperature and concentration in this so-called shear-induced state (SIS).
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 63

0.01

y (lis) =0

Figure 3.18 SANS-patterns and flow behaviour of a 2.4 mmol/l solution of an equimolar
mixture oftetradecyltrimethylammoniumbromide and sodiumsalicylate in heavy
water CD20) at T =20 oc, from Heen (1993)
64 CHAPTER II!

In addition, also the geometry of the viscometer, e.g. the slit width of a Couette viscometer,
is an influence. At high shear rates the shear induced state vanishes and a second
Newtonian region in the shear viscosity behaviour seems to occur. In the SIS the surfactant
solutions become viscoelastic with different time constants involved for the formation and
the relaxation of the SIS. This is elucidated in Fig. 3.17.
Additional information on the behaviour of the structure of the rod-like micelles can be
obtained by flow birefringence and scattering methods, such as light scattering (LS) or
small-angle-neutron scattering (SANS). The alignment of the rods due to shear causes the
refractive index to become ansiotropic, and an anisotropic scattering of light and neutrons
can be observed. By using these methods it has been shown that the micelles are nearly
completely aligned in flow direction in the SIS. Fig. 3.18 provides such SANS-results.
In the SIS larger, superordered structures have to be built up, because the shear viscosity
increases when the rod-like micelles are aligned in flow direction. A new shear induced
phase, like in the shear induced crystallization in the flow of polymer melts, with a
superordered, spatial arrangement of the rod-like micelles, has to occur. However, the
geometry of the spatial arrangement is still unknown. The superordered structures has to be
much longer than the length of the original rod-like micelles, because the measured shear
viscosity in the SIS increases with the gap width of the Couette viscometer used. In
suspension rheology (see Chapter 3.6) this phenomenon occurs when the diameter of the
rigid spheres of a suspension becomes larger than a tenth of the gap width. Thus, a raw
estimate of the size of the shear induced structures results in sizes of 50 Jlm, or even larger.
(The length of a micelle at rest is 1 "' 250-500 nm, the smallest gap of the Couette
viscometer is about 100 Jlm. Thus, the viscosity is thought to depend on the gap width at
1/10 of the gap diameter, the superstructure would have to be of 10 jlm or 20 times larger
than the micelles at rest.) Two models for the spatial arrangement of the rod-like micelles in
the SIS were discussed, the formation of pseudo-lattices from individual rods with lattice
planes sliding past one another, caused by cooperative electrostatic interaction energies and
the coalescence of the rod-like micelles to very long flexible rods that act like high molecular
polymers, Rehage et al. (1986) and LObi et al. (1986). However, the available SANS and
light scattering results do not confirm one of these models. This may be due to a non-
uniform size distribution of the shear induced structures, which could also vary in time. The
shear stress and the first normal stress difference exhibit small variations with time in the
SIS.
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 65

3.6 Suspension rheology of solid particles

In many processes in industrial engineering suspensions with various physical properties


are found: (glass) fibres are added to polymer materials in order to improve their material
properties or to reduce costs. In the production of paste paints for printing as well as in the
production of paper one has to deal with suspensions. For an understanding of the physical
processes involved in hydrotransport the flow behaviour of suspensions at high shear rates
is of importance. Also our human blood is a suspension, because it contains solid particles,
such as the red cells. This short review elucidates the importance of suspension rheology
for many processes. Nevertheless, it is not the aim of this chapter to deliver a complete
review on suspension rheology for all fields of application. For this purpose the reader
should refer to review articles, e.g. Barnes (1981), or the various textbooks on rheology.
The following deals with the rheology of those suspensions which can exhibit drag
reduction in the turbulent flow regime. The parameters influencing the rheology of
suspensions will be presented and it will be shown in which way they influence the
rheological behaviour. First suspensions of rigid spherical particles will be considered, and
afterwards suspensions with anisotropic particles, e.g. fibres.
The flow behaviour of suspensions is strongly influenced by the phase volume , c!>. of the
suspended solid particles, which is that fraction of space which is occupied by the suspen-
ded material. Dilute and semi-dilute suspensions of rigid spherical particles behave
Newtonian-like. The viscosity of a dilute suspension of rigid spherical particles was given
by Einstein (1906)

1Js = TJ(l + 2.5c!>) (3.47)

T} is the viscosity of the suspending medium. This equation holds for phase volumes up to
10 %. Note, that in eq. (3.47) there is no influence of the particle size. Furthermore,
homogeneous suspensions are considered, i.e. sedimentation effects are excluded. The
phase volume has a strong influence on the flow behaviour of suspensions, because
hydrodynamic forces act on the surface of the particles or aggregates of particles. Hence,
the density of the suspended material is of no influence. The viscous forces are proportional
to the velocity difference between the particles and the surrounding fluid. In other words,
particles exhibit inertial effects on the fluid motion. Batchelor (1977) extended eq. (3.47)
for semi-dilute suspensions of rigid spherical particles, (c!> < 0.15), by considering the
presence of other particles, i.e. interaction of the particles to
66 CHAP'IERIII

(3.48)

In flowing suspensions the flow behaviour can be influenced by additional colloidal forces
between the suspended particles. In general, there are the van-der-Waals forces, which
cause an attraction of the particles, forces due to absorbed molecules on the surface of the
particles, which cause a steric repulsion, and electrostatic forces, which also cause a
repulsion when the charge on the particles is of the same sign. Each of these forces strongly
depends on the distance between the particles. Thus the estimation of the overall colloidal
force acting on the particles can be quite complex. When these colloidal forces are larger
than those of the Brownian motion, the sum of these forces, which strongly depends on the
distance between the particles and thus on the phase volume, <)l, is responsible for the
formation of an internal structure. Aggregates can be formed if the attraction forces
dominate, whereas, vice-versa, pseudo-lattices are formed when the repulsion forces
dominate. These structures can be modified when surfactants are added to the suspension.
The electrostatic interaction can be influenced by additionally dissolving electrolytes. If
counterions are adsorbed on the particles, the electrostatic potential, which causes a
repulsion between the particles, will be weakened.
Aggregates can be destroyed in shear and, even more efficiently, in elongational flows. In
concentrated suspensions also flow induced structures can be built up in the suspension
under shear. However, since we are restricting ourselves to the rheology of drag reducing
fluids and drag reducing suspensions in the dilute or semi-dilute concentration range, this
effect will not be considered. Suspensions of particles of asymmetric shape, such as rods,
ellipsoids or fibres, are of great importance for drag reduction.
So far spherical particles were considered. However, when the shape of the particles differs
from that of a sphere, the viscosity increases at identical volume fractions. The largest
deviation from that of spherical-shaped particles occurs when the particles are of rod-like
shape. However, this fibre geometry is, as we will see later on, the most important particle
geometry which can produce drag reduction in turbulent flows when added to a Newtonian
fluid.
Furthermore, the ratio of the length to the diameter of a fibre, (LID) ratio, is the relevant
parameter for the tendency of a suspension to form temporary networks, Nicodemo &
Nicolais (1974). When the average size of the network structure depends on the shear rate
in a simple shear flow, the suspension will exhibit a pronounced shear thinning behaviour.
This can even occur at quite low volume fractions, which is similar to the behaviour of
surfactant solutions containing rod-like micelles, see chap. 3.5.
In some cases a threshold value of the shear rate or shear stress is necessary for the rupture
of those fibre networks. This can result in internal slip planes in a shear flow. At low fibre
RHEOLOGY OF POLYMER AND SURFACTANT SOLUTIONS 67

concentrations in laminar flows fibre plug flows are observed in these geomenies. Since the
strength of a fibre network is mainly influenced by the longest fibres, it is not the average
fibre length which is responsible for the network effects. The fibre length disnibution is of
importance. It may also happen that it is changed, due to degradation effects in a flow
situation, Maschmeyer & Hill (1974).
In elongational flows the fibres become aligned with their long axis in flow direction.
Consequently, the resistance against deformation becomes its maximum value. Thus in fibre
suspensions the elongational viscosity in a uniaxial extensional flow is always larger than
three times the shear viscosity, as is the case for Newtonian fluids. This has been
experimentally observed even in dilute suspensions. Theoretically it was shown for non-
dilute fibre suspensions by Batchelor (1971).
In shear flows of fibre suspensions at low shear rates the fibres rotate in the flow with an
approximately constant rotational speed. However, with increasing shear rate this rotation
becomes non-uniform. The residence time of the fibres in some directions becomes larger
than in other directions, i.e. the fibres do flip-flop motions. The reason for this behaviour is
due to two competing effects: the shear flow and, on the other hand, the Brownian motion.
At low shear rates there is only a weak orientation of 450 of the long axis of the fibre to the
flow direction preferred. With increasing shear rate the fibres become more and more
aligned in flow direction. The alignment process is accompanied by a decrease in shear
viscosity of the suspension, because the hydrodynamic resistance of the fibres reaches a
minimum, see also chap. 3.5.
With the resnictions mentioned above the fibres can freely rotate in a shear flow, when their
concentration is dilute, i.e. the rotational volumes of the individual fibres do not overlap.

(3.49)

n is the number of fibres per volume. For fibres with increasing aspect ratio the limit
between a dilute and a semi-dilute suspension is shifted to lower volume fractions. When
the number of rods per volume is in the range of

(3.50)

the suspension is called semi-dilute. In this equation D means the diameter of the fibre
whereas L stands for its length. In shear flows of semi-dilute fibre suspensions the
probability of collisions between individual fibres is still quite low, but the fibres are
hindered in their rotation. As a consequence, the rotation is slowed down and coupled to a
translation. An explanation for rod-like polymers is given by Doi & Edwards (1978).
68 CHAPTER III

Due to the alignment of the axisymmetric particles in shear flows elastic material properties
in the suspension occur. They manifest themselves in shear flows in non-vanishing normal
stress differences, which are enhanced when the fibres form networks. Thus the normal
stresses are created only by entropy elasticity, when no fibre network is present. The
probability to form networks increases with the LID ratio of the fibres, the fibre
concentration and the flexibility of the fibres. In fibre networks entanglements and knots of
the fibres exist. As a consequence, elastic properties of a fibre network can occur in flow
situations in which no alignment of the fibres is observed. The formation of networks in
fibre suspensions at rest can be due to colloidal forces.
IV DRAG REDUCTION AND TURBULENCE

4.1 Introductionary notes

The subject of turbulence is by far too complicated to be handled in an introductionary


chapter. Therefore, the goal of these notes is to introduce those elements of a description
of turbulence which are relevant for representing the interaction of a turbulent flow field
with dissolved or suspended additives. The intention is to understand the main elements
of the drag reducing mechanism, with hope that in doing so we would also start to
understand more about the turbulent mechanism itself.
Since we restrict ourselves to systems of dilute solution or suspension, the interaction of
the additives with the turbulent flow will be described in the frame of a Newtonian fluid
The discussion of possible change in rheology and its consequences on the effect will be
postponed and treated in the following chapters where drag reduction by specific
additives is discussed. In addition, we follow the most common representation of
turbulence since most of the experimental results of drag reduction are given in this form.
However, since it is thought that the effect cannot be described in these terms, some more
recent ideas for describing aspects of a turbulent flow will be added.
One of the main manifestations of the interaction of the flow with the additives is the
change in the energy balance. Therefore, it is obvious that drag reduction should be
discussed first on the basis of the energy balance, and one of the main questions to be
discussed is whether the additives interact with the production or the dissipation
mechanisms of the turbulent flow. Although additives are able to interfere with both
mechanisms and their interaction can reduce the drag, this is not always the case. To
answer questions like these, the basic equation of a turbulent flow has to be discussed
with respect to the physical processes involved.
The differences between flows with and without additives are given in the relevant mean
value of flow properties, like the velocity distribution, as well as in all statistical
parameters of the turbulent flow. Therefore, a short review of the statistical description of
turbulence will be given together with the relevant space and time scales involved. This
description provides the basis for the discussion of possible origins of the drag reducing
effect. Such a discussion examines, for example, whether the effect is due to the
interaction of single molecules or particles with the flow, or whether the particles act as
an ensemble, even though they do not interact with one another.
69
70 CHAPTER IV

4.1.1 Equations based on conservation laws

Physical processes are largely determined by conservation laws. In Eulerian form the
flow of a fluid is described as the motion of a continuum by its instantaneous change of
the quantities of state in an elementary volume at a given position.
The mass conservation is given by the continuity equation

(4.1)

which reduces for flows of incompressible liquids for which p, the density, is a constant
to

()ui = 0
()xi (4.2)

In other words, the velocity field is a solenoidal vector field. In eq. (4.2) there appears no
explicit time variation even when the flow is unsteady. The conservation holds for the
elementary volume itself.
The change of momentum based on the conservation of momentum for a continuum fluid
with viscous stress directly proportional to rate of strain is described by the Navier-
Stokes equation, which states that the acceleration following the motion of the fluid is
equal to the molecular force per unit mass and the body forces at this mass. The
molecular force is a surface force and reduces to a sum of stress gradients divided by the
density.
The molecular stress is conventionally divided into a scalar pressure, p, equal to 1/3 times
the sum of the three normal stress components, and a stress due to deformation and bulk
dilatation, which is a second order tensor, the deviatoric stress tensor with components,
O"ij The equation of momentum preservation can then be written

(4.3)

With the given definition of p the sum of the normal components of the stress due to
deformation must be zero

(4.4)
DRAG REDUCTION AND TURBULENCE 71

If one assumes that O"ij is approximately a linear function of the various components of
the velocity gradient for sufficiently small magnitude of those components and that the
deviatoric stress generated in an element of the fluid by a given velocity gradient is
independent of the orientation of the element, it follows that O"ij can be formulated as.

(4.5)

corresponding to the fact that under the symmetry assumption there only remain two
independent elastic modules, for details see, e.g. Batchelor (1967). However, it has
become common practice to introduce the relation

2
Jlz = --J.L
3 (4.6)

which satisfies the definition of p and eq. (4.4). This does not mean that the individual
normal-stress components are zero. This relation corresponds to a redefinition of the
pressure, which is no longer the thermodynamic pressure. Therefore, the statement that
the fluid is Newtonian means that Jl is independent of the rate of strain and that Jl2 is
assumed to be so too.
These relations are shown here in a certain length since the assumptions used in case of a
Newtonian fluid have to be compared later on with those of solutions containing very
long chain-like molecules which may exhibit some directional preference owing to
alignment of those molecules in a manner which depends on the past history of the
motion.
Additionally, the fact that in the Newtonian case the momentum equation can be written
by using a single coefficient for the viscosity veil the fact that also in this case the
extensional and the shear viscosity differ, however given by a universal relation the
Trouton ratio, TR, eq. (3.22)
For liquids consisting of long molecular chains the deviatoric stress may cease to be
accurate at only moderate rates of strain and little is known about how the expression
should be modified for such liquids.
In case of constant density and viscosity, the Navier-Stokes equation reduces to the form

dU dU dp (} 2u
+u--1 = - - + v - -1 +f1
ax.
1
--
at: Jax.J I
ax~
J (4.7)
72 CHAPTER IV

with the kinematic viscosity, V=Jl/p. For several applications it makes sense not to split
the molecular stresses into p and O"ij and to define the sum of them as

(4.8)

Another conservation equation is formulated by the first law of thermodynamics, which


states that the internal energy, E, in a closed system is constant.
The total rate of work done on a material element per unit mass of fluid is

(4.9)

The first term on the right of expression (4.9) describes the gain in kinetic energy of bulk
motion of the element. The second term is the work done by deforming the element
without changing its velocity. This work increases the internal energy. Assuming a fluid
of homogeneous heat content and neglecting a thermal heat conduction, the total rate of
change in the internal energy is given by the second term. This means

(4.10)

By substituting eq. (4.8) into eq. (4.10)

(4.11)

where the second term on the right is the rate of dissipation, <!>,of mechanical energy per
unit mass of fluid due to viscosity,

(4.12)

The total change in energy, which means the change in kinetic and inner energy (without
adding heat) per unit mass of fluid, is therefore
DRAG REDUCTION AND TIJRBULENCE 73

-D (1
-uu+E ) =u--+-+uf
Dui DE
Dt2 11 1 Dt nt 11
(4.13)

where the ftrst term on the left stands for the scalar product of the Navier-Stokes eq.
(4.3) withy, where the second term on the right contains the dissipation of mechanical
energy per unit mass on the fluid by an irreversible transfer into heat.
The net viscous forces per unit volume are

(4.14)

and they can be determined by the local gradients of the vorticity, !!l, of that element

(4.15)

with

(4.16)

Therefore, the net viscous force is significant only where the vorticity gradients are large.
The equation of motion of vorticity, obtained by taking the curl of both sides of the
equation of motion (4.7) with fi taken as a potential force is

aro. aro. au- a 2 ro


=-u--1 +ro-1 +v--1
J ax. J ax.
--1
at ax~
J J J (4.17a)

or

Dro ilu + v--~


= ro--~ () 2ro
ax.J
__I
nt J ax~
J (4.17b)

The first term on the right in eq. (4.17a) is the rate of change of vorticity due to
convection of fluid in which the vorticity is non-uniform past a certain point. The second
term represents the effect of velocity variations on the vorticity. The interpretation of this
term varies with the details of the velocity field, but it may be illustrated if mis regarded
74 CHAPTER IV

as a material line element instantaneously coinciding with a portion of the vortex line. A
part of the change in vorticity results from rigid body rotation of the line element, and
another part of a change in mresults from extension or contraction of the line element, the
so-called vortex stretching process. For an inviscid fluid it can be shown that a vortex
line always consists of the same fluid particles. The third term represents the rate of
change of mdue to molecular diffusion of vorticity.The diffusion of vorticity is related to
that of the momentum through that of the momentum gradient.
For an inviscid fluid a fluid of zero vorticity will continue to have zero vorticity. This is
no longer the case for a viscous fluid since diffusion of vorticity from a neighbouring
particle can occur. However, diffusion cannot create vorticity; vorticity can only originate
from boundaries. The solid boundary is a distributed source of vorticity. Together with
the stretching mechanism this is the main reason why the interaction of additives with the
flow is so important in the wall near zone where no-slip condition at the wall .

(4.18)

is a source of vorticity.
Due to this property the tangential vorticity source strength has a simple relation to the
pressure gradient, at least for a flow over a stationary plane surface.

(4.19)

The transfer by diffusion exactly balances the transfer by the pressure gradient.
Another conservation law holds between the first term on the right in eq. (4.17b) and the
change in in the circulation, r, defined by

(4.20)
The rate of change in circulation is represented by the rate of change of the flux of
vorticity across the material element, .Q.E. by

dor = Droi OF + ro doFi


dt Dt I I dt (4.21)

Changes in the magnitude and direction of BE have an effect on the flux of vorticity,
which is exactly balanced by the term of eq. (4.17b) mentioned. Eq. (4.22) can be
DRAG REDUCTION AND TURBULENCE 75

integrated over an arbitrary open material surface and, thereby, find the rate of change of
circulation around the boundary curve. From this it follows that

(4.22)

Therefore the flux of vorticity across a material surface element changes only as a
consequence of molecular diffusion.
The vorticity equation for Newtonian fluids, as we will see, is the equation which helps
us best to understand the energy transfer in a flow and is, therefore, of utmost importance
also for explaining drag reduction.
Before this equation is discussed in its decomposed form for turbulent flows, some
properties shall be mentioned in a more general form. If the flow is two-dimensional or
unidirectional, the vorticity is orthogonal to !l- Eq. (4.17b) reduces to

(4.23)

and takes the form of a diffusion equation. When the vorticity is parallel to the velocity
vector, then

(4.24)

defining for example the Beltrami flow for which the helicity density

(4.25)

determines the flow. The helicity as the integral property of a given domain, V, given by

H(t) =: JUj(J)jdV
v (4.26)

is for an in viscid fluid an invariant of the flow and, therefore, a conserved quantity. This
result was discovered by Moreau (1961) and rediscovered together with its topological
interpretation by Moffatt (1969). For details see Moffatt & Tsinober (1992).
76 CHAPTER IV

4.1.2. Reynolds decomposition

The equations based on conservation laws describe the dynamical behaviour of a


Newtonian fluid in all its details. However, their nonlinear form does not allow to solve
them in a closed form. Moreover, we know that turbulent flows are characterized by
velocity fluctuations. Therefore, all physical properties derived from the velocity field
remain fluctuating too.
Reynolds suggested describing the behaviour of a turbulent flow by decomposing y and
pinto

-u =il+u'
- - (4.27)
p = p+p' (4.28)

where ii and pare the time means of the velocity and of the pressure, respectively, for an
imagined infinite time and fixed boundary conditions, and .!L and p' are their turbulent
fluctuations. This is the simplest possible splitting and it is not as arbitrary as one might
feel at first because it leads to self-consistent equations with useful physical
interpretations. This has to do with the fact that for a given flow the mean is an
asymptotic state which only depends on the boundary conditions of the flow.
Nevertheless, this splitting is artificial.
All propositions for splitting y_still follow this line by splitting into a time mean and, for
example, into a phase locked average

(4.29)

where ii. is thought to be an oscillatory part of the velocity field. y' is the remaining
turbulent fluctuation no longer identical, however, with the one given by eq. (4.30), (see
Hussain & Reynolds, 1970).
With the new concept that the turbulent flow consists of typical forms of motion, so-
called "coherent structures", the splitting can be based on conditional sampling methods
or pattern recognition procedures.
If a certain splitting was chosen, normally the one given by eq.(4.27 &28), in all
equations in which y and p appear in section 4.l.l,y and p have to be replaced by their
mean and their fluctuations. It is obvious that eq. (4.2 & 3) would now consist of four
additional unknowns, the fluctuations of the variables. It is one of the old topics in
turbulence to find additional equations to close the system of equations. It has been
DRAG REDUCTION AND TURBULENCE 77

recognized that even in the time average this is a problem not yet solved- if it can be
solved at all.

4.1.3. The flow equations in turbulent decomposition

By definition, u' and p' are zero in the time mean; the fluctuations appear in the time
mean equations only as products.
The equation of continuity (4.2) therefore becomes

(4.30)

and the Navier-Stokes equation (4.3) becomes the Reynolds equation

(4.31)
with

:!:ij = -p8ij + 2J1Dij (4.32)


and

(4.33)

The contribution of the turbulent motion to the mean stress tensor is, therefore, given by
the so-called Reynolds stress tensor

-pujuj (4.34)

By formulating an overall stress tensor as

TIJ = 1::IJ - pu!u~


I J (4.35)

eq. (4.31) can be written as


78 CHAPTER IV

ii. aui = ~(Tij)


Jax.J ax.J p
(4.36)

By taking the divergence of eq. (4.31) and using the continuity equation, the pressure is
determined by the velocity as a formal Poisson solution by

_
p+p, =1-
41t
J{dx:axa
I
2
- - - [(-
Uj+Uj')(-
J J
dV(!)
U+U')]} -
J
-
lx'-
- -xl
v (4.37)

with x' being a position at which the pressure is to be found. From this equation it
follows that the pressure is not a local quantity but depends on an integral over an entire
velocity field.
The equation of the kinetic energy of the mean flow and of the turbulence is the one
which has to be discussed in the context of drag reduction. The kinetic energy of the
mean flow, 1/2 pujUj, is obtained by multiplying eq. (4.36) by Uj. The energy equation
becomes

(4.38)

The first term on the right in eq. (4.38) represents the spatial energy transfer of the mean
flow energy by the stress, Tij The second is the deformation work. It represents the
kinetic energy of the mean flow that is lost to or retrieved from the agency that generates
the stress. By substituting eq. (4.32) into eq. (4.35) and this into eq. (4.38), the energy
equation for the mean flow becomes

-U a-(-UU
1 - - ) =a
Jdx211
- (- -P-
pJ
dx
-n .. -UUU
u + 2 VU -,-;-) -
IIJ ljl
2V D ....
D +UU
IJIJ
-,-;D ..
ljlj
J J (4.39)

The last term is the turbulence production term, whereas the third term on the right in the
brackets is the transport term caused by turbulent motion. With this equation (4.39) for
the energy of the mean flow we did not obtain more information than with the momentum
equation for the mean flow. The equation for the turbulent kinetic energy is given by the
so-called turbulent energy budget or energy- transport equation and reads
DRAG REDUCTION AND TURBULENCE 79

_U a- (-UU
1---,---,) = -a- (-up+-UUU-2Vud
1-, 1-,-,-, -,-,-) -UUD-2Vdd
---,---, - ,,
J dx . 2 I I dx . p J 2 I I J I IJ I J IJ IJ IJ
J J (4.40)
with
, 1(iluf
d=- - +iluj)
-
IJ 2 dxJ dxI
(4.41)

The turbulence production , P,

(4.42)

has opposite signs in eq. (4.39) and eq.(4.40). Thus, this term serves to exchange kinetic
energy between the mean flow and the turbulence. A characteristic feature of a 3-D
turbulent flow is that it is of dissipative nature, which means that this energy exchange
normally involves a loss of energy of the mean flow to the turbulence.
The last term in eq. (4.40) is quadratic in d'ij. and, therefore, there is always a drain of
energy. It represents the deformation work viscous stresses perform against the
fluctuating rate. The term is therefore called the viscous dissipation, e,

E = 2vdfjdlj (4.43)

It is essential to the dynamics of turbulence since nonnally

(4.44)

At high Reynolds numbers the strain rate, d'ij. is much larger than the mean rate of
strain, Dij. This implies that the eddies contributing most to the dissipation of energy
have very small convective time scales compared to the time scale of the flow, and
therefore little direct interaction exists between the strain rate fluctuation and the mean
flow. The small scale structure of turbulence tends to be independent of any orientation
effect introduced by the mean shear. The drag reducing additives are of small scale and
they introduce an orientation in the fluid. This will be essential when explaining the
effect.
The production term, P, does not generally balance dissipation (integrally it does).
However, in most shear flows this is approximately the case

(4.45)
80 CHAPTER IV

Additional insights into turbulence dynamics is achieved by studying the vorticity of their
related fields.
If for turbulent flows the vorticity is split in the same way as the velocity or the pressure,
eq. (4.27 & 28), into

roi = ooi + roj with roj = 0 (4.46)

and both terms on the right hand side in eq. (4.46) are solenoidal, then eq. (4.17b)
becomes in the time mean

~- ~ ~2-
-Uaroi , uroi -,-, - D u roi
- - = -U--+ffid .. + ffi .. + V - -
J Ox j J Ox j J IJ J IJ OXJ
(4.47)

The first term on the right in eq. (4.47) can be written as

(4.48)

and gives the mean transport of ro'i through its interaction with the velocity fluctuations,
U'j, in the direction Of the gradient, d/dXj-
The second term is the gain or loss of mean vorticity caused by the stretching and the
rotation of the fluctuating vorticity component caused, in tum, by the fluctuating strain
rate.
Eq. (4.47) shows how velocity interacts with vorticity. However, if one is interested in
the direct interaction of vorticity with vorticity, one has to discuss the equation for the
square of the mean vorticity. Eq. (4.47) multiplied by ffii becomes

_ 0
U--
J
(1--)
2 I I =---
Ox . -ffiffi
o (-I -,-,)
Ox . ffiffiU
I J
---,--,oooi - -
+Uffi--+ffiffi
J I OX . l J
D --,,
"+ffiffid .. +
IJ I J IJ
J J J

v-- 2
(1- _) - -aooi
a 2 -mimi - vaooi --
dxJ 2 dxJ OxJ
(4.49)

and for the mean square vorticity fluctuations


DRAG REDUCTION AND TURBULENCE 81

_U -()- ( -CO
1--,) -,--; ()roi 1 d ,....,.....-; , 'd' --;-;D
CO = -U CO--- - - - U COCO +CO CO + COCO
J dx 2 I I J I dx 2 dx J I I I J IJ 1 J IJ
J J J

_ --,--, () 2 ( 1 --;-;) dcof dcof


COCOd + V - - -COCO - V - - - -
J I IJ dx~ 2 I I dx dx
J J J (4.50)

The importance of eq. (4.50) stems from the fact that the enstrophy

co!co!
I I (4.51)

and the dissipation, E, eq. (4.43), are discerned only via a divergence term

()2......,....,
Uj U j
- ,- ,
IJ IJ I I
---,--;
dd- COCO =-
dx- -
dx.
J I (4.52)
or
-- () ufu~ -- 2 () ' 2
= vco!co! + 2v---J = vco!co! + 2v_P_
I I dxdx I I dx~
J I I (4.53)

For high Reynolds numbers the last term in eq. (4.53) can be neglected and

=vcofroi (4.54)

Eq. (4.53) is the turbulent counterpart to the relations of eq. (4.14 & 4.15) for the general
flow field. An interesting result which follows from eq. (4.53) is that the dissipation for a
flow with the symmetry properties, oui/OXj= OUyOXj, as they exist for a potential flow,
can be given as

-- --2-
d2 u!u!
~2---
() u!1 u uu
= 2v-- 1 - 1 - 2u!-- = v--J_J
dx Jdx"J I dx~ dx~I
I (4.55)

which means that the energy dissipation reduces to a surface integral

/EdV
V
= JdujdO 2
dA

v (4.56)
82 CHAPTER IV

This equation shows that finite dissipation is still possible due to finite contributions from
a vortex sheet on the surface, A.
The only amplification mechanism of vorticity in the flow is due to the interaction of
vorticity with the rate of strain tensor. It is thought that due to this interaction the
stretching of the vorticity filaments strongly exceeds their compression. The consequence
of this somewhat irreversible tendency for the ens trophy balance in the vorticity equation
when

1
D-roro
2 I I ::.2
U (J)
---'"'----
Dt
= IDID d .. + VID - -
I J IJ I Ox~
1

J (4.57)

is that the first term on the right in eq. (4.57), which is called the enstrophy generation
term, is mostly positive and thus enables the flow to maintain the enstrophy.
Kinetic energy passes from the mean flow down through vortex motion of an ever
smaller length scale until it is finally dissipated in terms of viscosity. This energy transfer
process is called the energy cascade and is independent of viscosity until the final stage is
reached. Viscosity causes dissipation but does not control its rate. The intensity and
length scale of the small-scale motion adjust themselves so as to dissipate all the energy
transferred from larger scales.

4.2 Additional statistical means of turbulence

4.2.l.Correlation as definitions of scales

In chapters 4.1.1 - 3. the equations of flow at single points were formulated. We know,
however, that the motions of the fluid at two neighbouring points are not independent. In
a turbulent flow, flow structures of different scales coexist. To investigate the statistics of
motion of different spatial scales, we need information at different locations in the flow
field, at least at two points. It is, therefore, obvious that in this chapter only a very
introductive text can be given. For more detailed information see e.g. Tennekes &
Lumley (1972), Frost & Moulden (1977) and Panchev (1971). To get an idea of the
potential of statistical methods and the spectral decompositions of velocity measurements,
a short discussion of the most simple case of interaction of the flow field at two points
will be discussed briefly.
Assuming that the two locations at which the velocity is known are given by their
position vectors, x and x. which have a distance, lrl, then
DRAG REDUCTION AND TURBULENCE 83

(4.58)

In this case the correlation coefficient of the space correlation of the velocity fluctuations
is defmed as

(4.59)

These space correlations can be used to interprete the scale and structure of a turbulent
motion.
The most frequently used correlations are those between the same velocity components at
points separated in a direction either parallel (longitudinal) or perpendicular (lateral) to the
direction of the velocity, Rii, for which Rii(r=O)= 1 holds. For large values of r the
velocity fluctuations become independent of one another and R(r~oo)= 0. The
consequence is a correlation function, as shown in Fig. 4.1, with a maximum of R= 1 at
r= 0. Consequently, while at r= 0 dR/dr also vanishes, the curvature does not and instead
becomes large. A negative region in the correlation curve implies that u'i and u'i are in
opposite directions rather than in the same direction.

R (r)
XX r

r r

Figure 4.1 Typical correlation curves of the same velocity fluctuation components. Tbe
upper in longitudinal direction, the lower in lateral direction

The perpendicular correlation should normally be of the two opposed forms, as sketched
in Fig. 4.1, because continuity requires the instantaneous transport of fluid across any
plane to be zero.
A correlation function indicates the distance over which the motion at one point
significantly affects the motiom at another point. Such information can be used to assign
a length scale to the turbulent structure.
84 CHAPTER IV

Besides this space correlation a time correlation can be defined for the same velocity
components at a single point at different instants. For a single component, Rii(t) is called
autocorrelation. This correlation depends on t, the time separation, in a similar way as
the space correlation depends on r.

(4.60)
with

t=t+t (4.61)

If the velocity of the mean flow is large, then the turbulence structure is advected so fast
that at a point of observation the fluctuating pattern practically does not change and the
autocorrelation will be directly related to the corresponding space correlation by

r
t=-
ii (4.62)

This transformation is called Taylor's hypothesis.

If, instead of a pure space correlation at different positions, the fluctuations are also
correlated at different instants, this is called space-time correlation. This kind of
correlation is used for discussing probable trajectories of structures rather than the scale
of the structure itself.
Length scales for turbulence can be arbitrarily defined from the geometrical properties of
the correlation (eq. (4.59)). The most common are the integral scales:

Aij = JRijdr
0 (4.63)

The best known is the longitudinal length scale

A 11 = JR 11 dr
0 (4.64)

often also abbreviated as Ax or (Ln,Lx). This integral length scale is considerably bigger
than the lateral length scale
DRAG REDUCTION AND TIJRBULENCE 85

A 22 = JR 22 dr
0 (4.65)

for shear flows near a wall.


From the time correlation a micro-time scale is defined from eq. (4.60) as follows: The
curvature at the top of the correlation, Fig. 4.2, is determined by the small scales of the
turbulence. Expanding the correlation in a Taylor's series gives

R('t)=l+- aRI
1 -2
1:+-- 1:2 + ... a2RI
(),; -r=O 2 (),; -r=O
(4.66)

Owing to symmetry, dR/dt"'O, eq.(4.66), at small values oft reduces to

(4.67)
where

A =-~!_2 ()2R
't - a,;2
(4.68)

is called micro-time scale of the turbulence. This means this scale can be found by the
interception of the R(t)={} axis by the parabolic curve filled by R(t) at t=O, Fig. 4.2.

R(t)
1
~
',~~
.
."'
\

\"'-
\
\ "-..
\
\
'-.....
............. __
o~------._----------~----~
0

Figure 4.2 Sketch of an autocorrelation coefficient and the scales defined by this
function
86 CHAPTER IV

It can be shown that the micro-scale is proportional to the ratio

(4.69)
and therefore this is a measure of the most rapid change that occurs in the fluctuations of
u' i(t).
Analogical to the time correlation a micro-length scale or Taylor micro scale can be
defined as

r=O (4.70)

provided Rij is symmetric near r=O. The most frequently used micro scales are the
longitudinal micro scale

(4.71)

and the lateral micro scale

(4.72)

In practice the length scales are normally determined by the application of Taylor's
hypothesis.
The Taylor micro-length scale is not the smallest scale of turbulence. The dissipation
length scale (Kolmogorov scale),~. is smaller. Kolmogorov showed that the smallest
scale of turbulent motion must be related to the energy dissipation, eq. (4.43), and the
kinematic viscosity by

(4.73)
DRAG REDUCTION AND 11JRBULENCE 87

in terms of a time scale, the Kolmogorov time scale, this is

(4.74)

To characterize deviation from a nonnal distribution the higher moments of the flow field
have to be investigated.
The mean square departure, cr2, from the mean value, u, is called the variance or second
moment. It is defined by

(4.75)

If P(u') lacks symmetry with respect to the median, this can be shown by the third
moment, defined by

(4.76)

Its normalized form is called the skewness, S

(4.77)

The fourth moment , normalized by a4, is called kurtosis or flatness factor, K

(4.78)

It characterizes the symmetric deviation from the normal probability function, for which
K is 3. The flatness is large if P(u') has relatively large values in the tails.
These moments can also be defined for the velocity gradients. As Batchelor (1953)
showed, the skewness factor of the velocity gradient appears in the equation of the mean
square vorticity of the turbulence, eq. (4.50), as a factor in the term describing the rate
88 CHAPTER IV

mean of production of vorticity by stretching the vortex lines. For one component, for
example,

(4.79)
with

s, ~(::r {(::rr (4.80)

If S was zero, turbulent vorticity would be rapidly destroyed by viscous effects and the
high rate of energy dissipation could not be maintained.
Additionally, two statistical expressions shall be mentioned since they are of importance
in the discussion of drag reduction: the space-time correlation , a correlation like eq.
(4.59) but with an additional time shift as used in eq.(4.60), and the joint probabilities,
by which one examines, for example, the probability-density to find the value of u'iU'j in
an area element in the plane, (u'j,U'j). These joint probabilities are a measure of the
isotropy of the turbulence only, whereas the spectra of the space-time correlations, e.g.
Morrison et al. ( 1971 ), give the deviation of the velocity of a structure in relation to the
mean velocity (a wave speed).

4.2.2 Spectral description of turbulence

The correlations allow to judge the scales of the turbulent flow composed of different
scales of motion. To analyze them, it is convenient to use their three-dimensional Fourier
transform., or some other decomposition, e.g., a variety of wavelet representation, etc.
The reader is referred to descriptions for the spectral methods to Monin & Yaglom
(1971). The same applies to the analyzing techniques.
The transformed variable in space is the three-dimensional wave nwnber vector, kj. With
this the wave number spectral density for space correlations, eq. (4.59), is defined as

clt(k)
IJ -
=- 1 -Ju!ii~ exp(-ikr)dk
( 2 1t)3 I J -- -

= ~ IJJujiijexp(-ik 1r1 )exp(-ik 2r2 )exp( -ik3r3 )dr1dr2dr3


(2x) v (4.81)
DRAG REDUCTION AND TURBULENCE 89

We can rigorously regard <l>ij<.k)dlqdk2dk3 as the contribution of the elementary volume,


dk1dk2dk3. This is consistent with the inverse transform

ujiij = Jfllij(~) exp(i~!:)d~


vk (4.82)

which for r=O becomes

u!u~
I J
= J(k)dk
IJ- -
vk (4.83)

Frequently, the non-normalized form of R(r), eq. (4.59), is transformed in one


dimension only in form of a one-dimensional wave number spectrum

(4.84)
whose inverse is

+oo
ufuj(r1) = I wij(k 1)exp(ik 1r1)dk 1
(4.85)

The diagonal sum of the tensor, <l>ij. in eq. (4.81)

(4.86)

represents the kinetic energy at a given wave number vector. For r=O we find

+oo
2
I I = 3u' = JJJ<I>(k)dk
u!u! II - -
(4:87)

The directional information in <l>ij(k> is removed by integration over a spherical shell of


radius, k, (the modulus of k). With the surface element of the shell, dcr, we can write the
power spectrum , E(k)

(4.88)
with
90 CHAPTER IV

+JooE(k)dk=-UU
1__,----, =-U
3 ,2
-oo 2112 (4.89)

Analogically, the autocorrelation must be the Fourier transform of a continous,


symmetric, positive real function, S(ro), whose integral is unity. The Fourier transform,
S(ro) of R(t), is known as the power spectral density or simply spectrum.
It is defined by

+oo
R('t) = JS(ro)exp(i'tro)dro
(4.90)
whose inverse is

+oo
S(ro) = JR('t)exp(-i'tro)d't
(4.91)

This kind of description allows the introduction of some ideas about the interaction of the
structures of different scales. Vortex stretching involves an exchange of energy because
the strain rate performs deformation work on the vortices which are stretched. The energy
gained by a disturbance with velocity components, u'i and u'j, in a strain field, d'ij. is -
u'iU'jd'ij. per unit mass and time. Larger eddies are performing a strain in the smaller
ones and can consequently transfer energy to them. The inverse energy transfer is three-
fold: eddies grow in size (1) by vorticity diffusion, (2) by pairing, and (3) by vortex
breakdown processes.
Eddies of a given size are stretched by larger eddies whose strain-rate field is constantly
changing in magnitude and direction. The smaller the eddy scale the smaller the decrease
in anisotropy so that at small scales the strain-rate field itself may be expected to be
isotropic in the mean, local isotropy.
The spectrum has its specific form shown in Fig. 4.3. In the part of the spectrum in
which local isotropy prevails, time scales are short compared to those of the mean flow.
Therefore, small eddies are always in approximate equilibrium with local conditions in
the mean flow. For this reason, the range of wave numbers exhibiting local isotropy is
called the equilibrium range. In this range the energy extraction from the mean flow is
very small. However, all energy is finally dissipated by viscosity, and, therefore, the
total amount of energy transfer must be equal to the dissipation rate, E. The cascade
terminates when the Kolmogorov micro-scale, eq. (4.73) is achieved, with the associated
Kolmogorov velocity , v
DRAG REDUCTION AND TURBULENCE 91

1
v = (ve)4 (4.92)

E(k) energy universal


containing equilibrium
eddies range

Figure 4.3 Spectral distribution of the energy of turbulence. Region of (a)


anisotropic eddies, (b) inertial subrange, (c) dissipation. (1) The
transformation of energy from the mean flow. (2) The transfer of energy
from large to smasH eddies. (3) The dissipation of kinetic energy into heat
energy

Batchelor (1953), for example, showed that the spectrum of dissipation D(k) is related as
follows with the energy spectrum eq. (4.88)

(4.93)

The dissipation is proportional to the square of the velocity gradient, and the dissipation,
eq. (4.43), becomes

e = jn(k)dk = 2v jk 2E(k)dk
0 0 (4.94)

On the other end of the spectrum, for the large scales the viscosity is irrelevant. The
energy feeding results from the mean strain rate and the energy is transferred to small
scales at the rate . Thus, the large scale part of the spectrum should be based on Dij and
. Since Dij is not universal, neither is this part of the spectrum.
Between these two extreme parts of the spectrum there exists a range for which we
hypothesize that the energy flux across the inertial subrange is conservative. This means
that the flux across each wave number is constant and is equal to . This range is called
the inertial subrange in which the power spectrum is given by
92 CHAPTER IV

2 5
E(k) = <X3k 3 (4.95)

with a= 1.5. However, since this inertial subrange only exists if the Reynolds number is
quite large, it is unlikely that we will encounter a fully developed inertial subrange in
laboratory flows.
The energy spectrum is continuous. However, if one thinks that the spectrum is
composed of eddies of discrete size, one finds that eddies corresponding to a wave
number, k, get their energy of the order kE(k), mainly from eddies with a wave number
0.38k whose strain rate s(0.38k)=:l/2s(k). The anisotropy produced by the larger eddies
is the important parameter; it is proportional to s(0.38k)/s(k). The energy flux, T,
becomes

2 3 5
-- - -
T(k) =a 3E2k3 (4.96)

This shows that isotropy and energy flux are strongly related and a manipulation of this
process must be very effective.

4.2.3 The concept of coherent structures

First, it has to be elucidated that the concept of coherent structures is by no means a


description of turbulence, which is in contradiction to the previous formulations of
turbulence. Already in chapters 4.2.1-2 we introduced notions like structures, scales or
vortices of different size, etc. The new word is "coherent," and it means that the
structures we are talking about are more deterministically describable than what is thought
to emanate from a nonlinear coupling, as given by the Navier-Stokes equation.
The notion was derived from mainly two-dimensional observations. A long time ago it
was recognized that the basic structure elements of the turbulent flow are eddies. If one
assumes that turbulence can be described by a distribution of eddies of a certain kind, this
would be a strong restriction helping to simplify the description of the complex
mechanism of turbulence. If, in addition, similarity laws between the scales of these
vortices existed, the field pattern of the flow would be describable by fractal geometries.
On the other hand, the number of features proposed as basic elements have drastically
increased in the last decade, and it is not unreasonable to speak of a zoo of coherent
structures. One of the most complete descriptions of the concept and its elements can be
found in Robinson (1991) and Kline & Robinson, and Robinson (1990), respectively.
DRAG REDUCTION AND TURBULENCE 93

Based on the vortical origin of the basic elements, in some literature the word "coherent"
is related to areas where high vorticity is present. Their definition is then given with
respect to this property in several variants. This seems rather risky as vorticity is one of
the most local quantities of the flow field and the coherent structure expresses the fact that
durable features are present in the flow. Therefore, it is probably better to think of
vortices which incorporate a large portion of potential flow such as in a potential vortex
flow.
The fact that vorticity agglomerates in certain areas of the turbulent flow field is one of the
unsolved problems of turbulence. The existence of such inhomogeneities in the vorticity
field, however, has been known for a long time and interpreted as the intermittent
character of the turbulent flow. The word "coherent" implies not that these features have a
periodic behaviour in time and space, but on the contrary that they appear randomly in
time and space. "Coherent," we are well aware, is, therefore wording rather than a well
defined concept. Nevertheless, due to this "concept" at least two new structures of the
complex flow could be separated in a more deterministic form.
(1) In the layer very close to the wall of a wall-bounded turbulent flow, at least in the
mean, counter rotating longitudinal vortices of a diameter of about 50 viscous
units and a length of the order of ten diameters are present. They disrupt after a
characteristic time, Tp, the so-called burst period.
(2) During this burst phenomenon two very characteristic local motions can be
observed: the ejections and the sweeps. The ejections transport slow near-wall
fluid into the bulk of the flow, whereas the sweeps transport high-speed outer
fluid towards the wall. It has become common to define them as events which
show up in the second or the fourth quadrant in the (u', v') plane, respectively.
The intermittent time between events of this kind can be statistically analyzed. In
this way the mean period, Tp, can be evaluated. To make clear that this is not a
period in the classical sense at all, a measured hystogram is shown in Fig. 4.4.
As a warning the reader is reminded that one has to differentiate carefully between
physically existing structures and structures constucted by taking long time means, even
when these means have been evaluated with conditional sampling methods. From Fig.
4.4 the artificiality of the burst period is evident and elucidates how problematic it is to
develop models based on this value.
The main input from this concept that helps to understand the drag reducing mechanism
has to do with our desire to investigate the behaviour of the additives in typical flow
fields. The hypothesis is that the main interaction of the additives with the turbulent flow
field could be understood by a statistic over the interaction in the selected flow features
94 CHAPTER IV

y+ 1ft
0 2.7 59
D 4.0 60

0.00=~------..l...r..;;;...-----.----.........--1
0 100 200 300

Figure 4.4 Histograms of the time, Tp, between two consecutive events of the same
class.

4.3 Fundamentals used to explain drag reduction

For each kind of additives the argumentation of how drag reduction is achieved can vary
and will be discussed accordingly. In this chapter ideas shall be formulated which allow
to categorize the different explanations and to help find probable common roots of the
effect.
Primarily, we observe that certain additives in a highly diluted form are capable of
reducing the energy consumption of a turbulent flow. This is unexpected since, by
adding particles, the dilute suspension or solution has an increased shear viscosity and, in
terms of a Newtonian fluid, should dissipate energy more rigorously than the solvent.
The additives must, therefore, interact qualitatively differently with the flow field.
Several mechanisms have been proposed, and they have to be classified with regard to
some basic concepts used in the explanation:

4.3.1 A rheological explanation

Assuming we knew the correct constitutive equation of the specific drag reducing
solutions in the frame of a continuum mechanical description, then the equation of motion
could be given. However such constitutive equations, at least for the relevant fluids, have
not been found, and if they were, they would probably be too complicated to describe a
turbulent flow. In other words, it is hard to believe that drag reduction based on such a
description will be explainable.
DRAG REDUCTION AND TURBULENCE 95

Nevertheless, efforts of this kind have been made by direct numerical simulations of the
equation of motion with specific constitutive equations. But this seems a premature task,
as Leal confirmed (private communication), and does not contribute to any real
understanding of the problem at hand. If one had a model for polymer molecules that was
known to be accurate for complex, time-dependent flows, and a simulation that was
kinematically reliable for turbulence, then it would be of some interest to see how a dilute
solution of polymers would modify turbulence. It would be more promising to apply
such a method for rod-like fibres, a task which, to our knowledge, has not yet been
undertaken. The flow would have to be extremely accurately represented because the
polymer response in dilute solutions is highly sensitive to the details of the flow. Most of
the constitutive equations used, based on molecular models, are reasonably well
documented for simple shear or relatively simple extensional flows, but there is no
evidence whatsoever that they can provide decent results when exposed to flows with
relatively rapid changes in magnitude (and other details) of the local velocity gradient
tensor as in turbulent flows. For example, there is really no evidence that such models
respond dynamically in an accurate way.
If the polymer encounters regions of stretching and relaxation in some sequence, for
example, the model must be accurate in the time scales and details of both relaxation and
stretching if it is to yield qualitatively reasonable results too. This means that, for
example, for a bead-spring chain molecular model both the spring-law and the friction
dependence on conformation must be accurate in detail, not just roughly correct.
Substantially more effort is required, based on the comparison of experiments and model
predictions in much simpler time-dependent flows, to establish a model for dilute
solutions that can be trusted to give reliable results for such types of simulations.
A certain simplification would be to assume that the solution is so dilute that the relevant
structures of the flow are still describable by a Newtonian fluid, and that the influence of
the additives is restricted to limited zones, i.e., where the stress field is very high. This
concept is like separating the problem into two steps. The flow field determines where the
additives are active. The interaction with the additives can then be calculated, assuming
that the local flow field is known. By statistics not yet known it is hoped that this
interaction can be incorporated in order to describe integral properties like the energy
dissipation of the flow. It is evident that these ideas are very similar to the ones
encountered in the concept of coherent structures. In the frame of such a description the
additives would be active where the flow field itself has its active spots. Let's call this
mechanism a spotty interaction. The relation to the well-known intermittency of the
turbulent flow field is evident.
In the most extreme form, however, one could postulate that the local interaction of the
additives with the flow structures depends on the concentration of the additives at that
96 CHAPTER IV

location. In other words, since the concentration of additives in a certain volume depends
on the time history of the flow field, an additional time dependence would enter into the
description, which in a statistical form would have to be incorporated. This kind of
description is already at the limits of a continuum concept, and if we assume that single
particles or a very small amount of them could already be active in these spots, the
continuum approach to the problem is no longer valid.
These few remarks make clear that an explanation of the drag reducing effect should
always incorporate the material properties of the solution. At several points in the text we
made and will make some marginal notes where change in the rheology could be of
importance. However, our knowledge of the rheological behaviour of the drag reducing
solutions is too small as to expect a description of the effect in such "terms" now and in
the near future.

4.3.2 Explanations based on change in the energy transfer in the turbulent


flow

Drag reduction can principally be explained by at least two different influences of the
additives: (a) by reducing the turbulent energy production, which means the additives are
capable of damping the withdrawal of energy from the mean flow into the turbulent
fluctuations (the last term in eq. (4.39)), or (b) by reducing the dissipation of energy into
heat (the second term from the right in eq. (4.39)).
The main change in turbulence observable in drag reducing flows is the anisotropy in u'
and v' and, therefore, a drastic change in the tensors d'ij as well as in u'iU'j In other
words, in the drag reducing case the turbulence production, P, given by eq. (4.42) as
well as the main dissipation term on the far right in eq. (4.40) must change considerably.
Already these comments show that there is no such simple explanation such as the
additives interact with the production or the dissipation. However, in all drag reducing
cases the flow remains turbulent. This means that in all cases the enstrophy generation

(4.97)

has to be suppressed, since it is the only mechanism in turbulent flows sustaining a high
level of turbulent dissipation, eq. (4.54).
In chapter 5.5.3 the implication of this statement will be discussed in some detail,
especially in regard to how a change in the anisotropy of the velocity fluctuations modify
the production and dissipation mechanisms of a turbulent flow. In addition, the power
DRAG REDUCTION AND TURBULENCE 97

spectra, eq. (4.88), are used to interpret some of the changes in the energy balance, due
to the additives.
The drag reducing mechanism is a complex one and, therefore, so is its explanation.

4.3.3 Explanations by scaling arguments

The attempt to formulate an explanation of the drag reducing effect based on change in the
material properties seems up to now to be unrealistic. The same arguments can be used
with respect to a model based on a change in the energy transfer. The reason for this is,
again, the same; the mechanism itself is too complex. An alternative to this approach is to
somehow revers the problem by asking which quantities are best attainable. These could
be the quantities by which one could start, inhope of constructing a simplified model. -
The answer is the scales are the candidate.- They can be rather easily measured and they
can be incorporated in models based on dimensional analysis.
To elucidate this idea, let's look at a characteristic signal of a component of the velocity
fluctuations in a Newtonian and in a drag reducing fluid, Fig. 4.5.

a b
u' +0.3
[In/s]

!o~ 1\
JtJ
,rW~
I~ \ ( (\ ) ~

\J\1/
0
~ v~ ~~I
J ~ \(V IV

-0.3
0 0.5 0 0.5
t [s]

Figure 4.5 A sequence of the velocity fluctuation, u', in a smooth square pipe
(D=150mm) flow in flow direction, (a) in a Newtonian fluid at Re = 9600,
and (b) for a drag reducing fluid in the same pipe at Re= 9600 at y+= 25

A characteristic of a component of the velocity fluctuation in Newtonian fluids is that their


gradient change is much more abrupt and change more often their sign.The change in
sign of the derivative of the velocity fluctuations can, in grosso modo, be interpreted as a
change in sign of the vorticity. In other words, the locations of "unsteadiness" are
98 CHAPTER IV

boundaries of counterrotating vortical systems, or are at the edges of the core of a


potential vortex, such as a Rankine vortex. Therefore, the scales defined by areas of the
same sign in the velocity gradients are a measure of the length scale of the vortices or at
least their cores. In Newtonian fluids this scale is shorter than in the drag reducing case.
The power spectra in the two fluids show a shift to lower frequencies in the drag
reducing case, which is compatible with these findings.
As mentioned at the beginning of this paragraph, the specific ideas on how drag reduction
is achieved in the different systems will be explained as necessary. On the other hand, we
have given a rather detailed presentation of the basics of a turbulent flow. The objective is
to enable the reader to formulate his own ideas in such forms or to analyse his own data
with respect to such presentations with the advantage that the autocorrelation of du/dt is
given by

du'(t)d~'(t) = 0 , 2 d 2 .R(t-t) = _0 ,2 d2 R(-r;)


dtdt dtdt d-.; 2 (4.98)

with the relation

(4.99)
The spectrum of the first derivative is proportional to ro2S(co).
The relations between the spatial and temporal spectra are simple only in the isotropic
assumption, which, however, is not fulfilled for the most relevant cases investigated. In
general, they are quite complicated and their description would occupy too much space.
So the reader interested is asked to consult the relevant literature.
What shall be reviewed here, though in a very short form, is the energy cascade of a
turbulent flow vital for drag reduction.
This mechanism is directly related to the vorticity cascade, which, as we know, is due to
a vortex stretching and decay mechanism. The stretching is due to the strain rate field in
which the vortex line discussed is embedded. On the basis of conservation of angular
momentum, vorticity is amplified in direction of a positive strain rate and attenuated in
direction of a negative strain rate.
Another aspect of this process is the interaction of the nearest wall structures with the one
in the bulk flow. Since vorticity is only produced at the wall and therefore has to be
transported into the bulk flow, this mechanism is the key to this process. It is debatable
whether the process is kept up by an instability mechanism of the longitudinal vortices or
DRAG REDUCTION AND TURBULENCE 99

by large structures producing the outer pattern so that the instability process is triggered
by these outer structures. We think that a discussion along these lines is not so fruitful.

However, we do consider the micromechanical aspects and finally their modelling very
helpful in understanding the energy transfer processes going on close to the wall. The
concept of "coherent" structures will be used in this sense to explain at least part of the
phenomenon because the scales of these structures are still large compared with the
additives. This fact allows to investigate the influence which particles have on
characteristic structures present in the flow. The goal is then to describe the flow by
statistics on such features with and without additives and their altered interactions.
V DRAG REDUCTION IN POLYMER SOLUTIONS
5.1 Integral aspects or gross flow behaviour
5.1.1 The characteristic parameters
5.1.1.1 Dilute solutions

As mentioned in the introduction, the effect manifests in a decrease of pressure loss over
a fixed length of a straight pipe compared with a flow of Newtonian fluids through the
same pipe section at the same Reynolds numbers for Reynolds numbers which are higher
than a certain critical value. This fact is equivalent to a reduction of the friction factor due
to the added polymers, see Fig. 1.2 & 1.3. In this gross flow behaviour of the friction
factor the main characteristics of drag reducing flows by polymers can already be
observed.
It appears that the friction factors of turbulent pipe flows due to drag reducing polymer
solutions are bound between the Prandtl-Karman law, eq. (1.12), and a maximum drag
reduction asymptote found empirically by Virk (1971) and given by eq. (1.13). Between
these limits the observed friction factors are approximately linear in Prandtl-Karman
coordinates, as shown in Fig. 1.3. These linear curves begin at the Prandtl-Karman curve
showing an onset behaviour. With increasing wall shear stresses given in log(Re fl/2) the
friction factors of these dilute solutions follow the Prandtl-Karman law as if they were
Newtonian fluids, but suddenly, for a given wall shear stress, they change to a non-
Newtonian fluid behaviour. This onset value together with a slope different from
Newtonian fluids therefore defines the friction dependence of the given fluid. Or,
formulated by comparison with the Newtonian fluids, instead of the slope they use the
slope increment, B, in comparison with the slope of the Newtonian fluid.
Naturally, this friction behaviour has its counterpart in the mean velocity profiles of such
drag reducing fluids as described in the introduction and shown in law of the wall
coordinates by Fig. 1.4. These velocity profiles consist, like Newtonian fluids, of a
viscous sublayer which is usually thicker than in the Newtonian case, and a core region
with a logarithmic profile with the same slope as the one in Newtonian flows, but shifted
by a value M3. The zone between is called the elastic zone since it seems that only in this
intermediate zone the polymers cause deviations from the Newtonian fluid. To the left
this zone is limited by an asymptotic profile called Virk's ultimate profile, which is given
by the empirical eq. (1.16), a logarithmic profile which is steeper than the Newtonian one
by a factor of about 5. The inner logarithmic velocity profile can intersect with this
ultimate profile or can have a kind of a buffer zone profile, which can but does not have
to start on this asymptotic profile. It has to be pointed out that this intersection point is a
101
102 CHAPTERV

retro onset value; it marks the wall distance at which the polymer cannot interact anymore
with the core flow.
The asymptotic friction and velocity behaviour are independent of the diameter of the
pipe, D, and the type of polymer used; they are thought to be the manifestation of some
universal law. This is of utmost importance because this law shows that by drag reducing
polymer additives one cannot manipulate the turbulent flow between the Prandtl-Karman
law and the laminar conditions, but that the manipulation is limited by a state which is a
key to understanding drag reduction and, even more, to obtaining insight into how
turbulent flow can be manipulated in general.
This result was rather unexpected since, by using additives, the flow becomes dependent
on parameters like the polymer concentration, c, and the chosen material, A, characterized
by its molecular weight, its length and its flexibility, to mention only the most important
ones. By dimensional analysis the dynamic behaviour cannot be reduced to three
variables as in a flow of Newtonian fluid, and in fact, the other parameters, like the onset
wall shear stress, the slope increment of the friction factor and the parallel shift of the
velocity profile depend on additional variables. The parameters for which the dependence
was investigated are c, A and the pipe diameter, D, where the D dependence is of special
interest since it shows that in addition to the usual length scale, y, the wall distance, and
the molecular length scale, I, the D dependence takes care of the physical fact that the
polymer seems to interact differently in the near-wall region and at the centre of the pipe.
The onset wall shear stress in the Prandtl-Karman coordinates for a given polymer-
solvent system is essentially independent of the pipe diameter, polymer concentration and
solvent viscosity. If we consider the onset in a f-Re plot, the onset is shifted towards
higher Re with increasing pipe diameter, and towards lower Re with increasing
concentrations. The dependence of the onset wall shear, ('tw)o, stress from the polymer
species used is given by the radius of gyration through the empirical law

2<n<3
(5.1)

with QT being an onset constant, characteristic of a polymer-solvent system. The power


n is 3 as is normal used, however, values are found between 2 and 3. QT must be
evaluated experimentally. For several systems it can be found in the literature. For the
two standard pairs, PEO-water QT is 4.42.110-21 Joule and for PAM-water QT is
13.15.110-21 Joule, (from Virk 1975).
The slope increment is also essentially independent of the pipe diameter, and varies
though with the concentration. Virk (1975) gives an approximate empirical relation by
DRAG REDUCTION IN POLYMER SOLUTIONS 103

which the increment, 8, in the dilute regime is proportional to the square root of the
concentration

(5.2)

This relation seems to be independent of the solvent. However, the polymer still has to
remain random-coiling in its most expanded conformation. With eq. (5.2) one can,
therefore, define an intrinsic slope increment, II, per macromolecule, suitable for further
correlation with macro-molecular parameters.

II= li~(M I c) (5.3)

As Virk (1975) showed, the intrinsic slope increment can be linked with the number of
monomers by

II= 1CN3/2 (5.4)

The main effect of polymer additives on the velocity profiles is the parallel shift of the
velocity profiles in the core region. Since the velocity profile in this part of the flow is
like that a Newtonian fluid, however, and starts at higher values, it is called the
Newtonian plug. It starts at a given wall distance. With increasing drag reduction this
boundary shifts towards the centreline of the pipe, which means that the Newtonian plug
shrinks. At maximum drag reduction this core disappears. In other words, maximum
drag reduction is achieved when the elastic zone (layer) pervades the entire cross section,
with the exception of a small ring at the wall where a viscous sublayer still exists, and
one zone very close to the centreline of the pipe where deviation from the logarithmic law
must exist. The asymptotic maximum drag reduction profile is thought to be due to a
change in the material properties of the fluid and with it the rheology in the whole interior
turbulent pipe flow. Between these limits, the profile shows both the elastic layer and a
Newtonian plug region. The maximum drag reduction profile has a mixing length
constant, K, of 0.085, which is about a fifth of the value, 1C= 0.4 for a Newtonian fluid.
It is important that the term "elastic layer" does not mean that the fluid in this zone is
differing from the Newtonian fluid mainly by an elastic term. It primarily means that in
this layer the molecules interact with the flow by deformation, and since the molecules
react elastically to strong deformations, this property can also be found for the fluid,
though the elastic behaviour can still be very locally distributed. A rheological description
would have to explain the effect based on molecular parameters such as the length, the
104 CHAPTER V

stiffness or the relaxation time of the molecules. Eqs. (5.3 & 5.4) already showed the
importance of the length of the molecules characterized by the molecular weight, M. It
should be mentioned that long, flexible molecules are more efficient than shorter or stiff
molecules. This point will be discussed later on in more detail since Virk & Wagger
(1990) pointed out that there is a great difference if the molecules are randomly coiled,
stretched or expanded.
In the framework of a discussion of the typical gross flow properties it has to be
mentioned that stiff macromolecules and, even more, surfactants and fibres can produce a
"retro-onset." This means that the typical velocity profile commenced on the maximum
drag reduction asymptote and then exhibited retro-onset into the polymeric regime of less
than maximum drag reduction. In Fig. 1.4 the retro-onset shows up in point A.

5.1.1.2 Semi-dilute solutions


10 1
'I Rhodopol 23
.
~

[Po sl 500ppm
Do
Do 1000ppm
2000ppm

'c..
'bo 0

.. .. ceo b

. ..
n

.......... ...
%_
.. ... .. ""'h,
. .. . .
.. ...... ,_
uoo
co
0

- 10 2 10 1

T [Pal

Figure 5.1:Shear viscosity versus shear stress for solutions of different concentrations of
Xanthan gum in water (Rhodopol23, Rhone Poulenc)

When in a polymer solution the polymer concentration is raised, the probability of


interactions between the individual polymer chains also increases. Finally, in a
concentrated polymer solution a polymer network exists which consists of entangled
polymer molecules. As shown in the previous chapters, in dilute polymer solutions the
shear thinning behaviour, see chapter 3, which is exhibited by almost all polymer
solutions has a negligible influence on the friction behaviour in turbulent flows.
Consequently, the onset-point of drag reduction was found to be independent of the
DRAG REDUCTION IN POLYMER SOLUTIONS 105

polymer concentration, and thus, the Reynolds number was calculated with the solvent
viscosity.
In polymer solutions in which higher concentrations are needed to produce drag reduction
in turbulent pipe flows, the influence of the shear thinning behaviour on the friction
behaviour of these semi-dilute solutions can no longer be neglected. This holds mainly
for solutions of relatively stiff rod-like polymer molecules, e.g. for Xanthan gum and

--
carboxymethylcellolose (CMC) solutions.
10- 2
Rhodopol 23 l
(a) ..SOOppm
- r--------..._ 1000ppm

. ------- t--- 2000ppm

. . .. .
a
f c t-- t---
. .. ...
A

t--
-. c

.. ... .
0

~~
0 ~t 0.

0 n

---1--------
u
c

1------1---
1--t--..
a o 10~

Re

p --- ....____
a
(b) 7


"'\ ------ ....__
. .. . ---......
---......
""'""" ~ . . -. .. ..
f

.
"~a ... ....
n '1.

K
Rhodopol 23
.. SOOppm
1000 ppm
2000ppm
a
~
3

Re

Figure 5.2:Friction behaviour of the Xanthan gum solutions of Fig. 5.1 in a 5.2 mm
pipe. In Figure a) the Reynolds number is based on the solution viscosity,
whereas in Figure b) the viscosity of the solution at the wall shear stress is
used for calculating the Reynolds number
106 CHAP'IERV

Fig. 5.1 provides an example of the shear viscosity behaviour of Xanthan gum solutions.
These solutions exhibit a remarkable shear thinning behaviour which becomes more
distinct with increasing polymer concentration.
As shown in Fig. 5.2 a) these Xanthan gum solutions exhibit drag reduction in a
turbulent pipe flow (d = 5.2 mm). For an engineering application this plot provides the
necessary information on the expected friction reduction due to adding the polymer. On
the other hand, if one is interested whether the friction behaviour of these solutions is
similar to the one found in dilute polymer solutions, one has to correct the Reynolds
number due to the influence of the increased solutions viscosity. Now the question arises
which viscosity of the shear thinning polymer solution should be used for this purpose.
If the shear viscosity for calculating the Reynolds number is taken for each wall shear
stress from Fig. 5.1, the friction behaviour shown in Fig. 5.2 b) looks quite different. At
the highest polymer concentration, c = 2000 ppm, the friction data now approach Virk's
maximum drag reduction asymptote which is the lower line in Fig. 5. These data
demonstrate that if the influence of the increased shear viscosity is considered, the friction
behaviour becomes similar to the one found for dilute polymer solutions. The data in Fig.
5.2 b) are a good example for the so-called retro-onset behaviour (see Chap. 5.5.4).
Semi-dilute poymer solutions of fle~ible polymer molecules, e.g. polyacrylamide, exhibit
a concentration dependent on onset behaviour. With increasing polymer concentration the
onset point will be shifted to lower Reynolds numbers which is in agreement with an
increase of the longest relaxation time of the polymer solutions. Here, it has to be
recalled, the limit between dilute and semi-dilute can occur at quite low polymer
concentrations, e.g. for polyacrylamide solutions of M = 410 6 g/mol at approximately
35 ppm.
The already encountered viscosity problem for dilute polymer solutions becomes even
more severe for semi-dilute solutions, due to the difficulties in finding the correct shear
stress in a turbulent flow at a certain wall distance. This is because the viscosity of shear
thinning polymer solutions can vary with the wall distance.

5.1.2 Up-scaling problem

There is a necessity to introduce an additional length scale parameter characterizing the


width of the elastic zone. One reason for this is the so called up-scaling problem. This
refers to the problem of predicting the drag reduction for a pipe of a different diameter
when we know the behaviour of a solution in a test pipe. In the literature several methods
can be found but all are to a high degree empirical solutions. To get an impression of
such a method, one of the most used, the two-loci method of Granville (1977, 1984), is
described. Since a theoretical description would only be possible if there existed, for
DRAG REDUCTION IN POLYMER SOLUTIONS 107

example, a theory for LlB and its relation to the amount of drag reduction involved in it,
the method is straight-forward and assumes a dependence between LlB and an additional
length scale parameter, 1e

(5.5)

20

I
~r

15

10~~--~~------~--~----~~----~------~.
fo 2 4 1o3 2 4
tn-.JfRe

Figure 5.3 Scheme of an up-scaling procedure by the method of two loci. In the vicinity
of a measured value (in the test pipe) in the friction factor plot in Prandtl-
Karman coordinates an upper and lower curve with given AB 1 and AB2 are
constructed using eq. (5.6). Through the measured value an interpolated
curve is drawn and intersected by the loci of the pipe diameter, for which
drag reduction shall be predicted using eq.(5.8). In the case shown, the up-
scaling is constructed for two pipe diameters one being two times larger and
the second, four times larger than the original test pipe diameter

The values of both AB, see Fig. 1.4, and le+, eq. (5. 7), have to be satisfied for pipes of
different diameters.
It is obvious that for a plot of 1/Vfagainst ...JfRe lines of constant LlB are parallel to each
other in accordance with

1 2.5
-=-lnv'fRe--1 ( 2.5Jn-Y8+-2.5-B
43 3 )
10 --B 2
v'f ..,fi ..,fi 30 ' 10 (5.6)
108 CHAPTERV

with B1,0 being the value ofB1 in the absence of the additives. Eq. (5.6) was found with
the assumption that the elastic velocity profile holds up to the wall, which means
neglecting any effect from the viscous sublayet and the buffer zone. Therefore, eq. (5.6)
is a first locus, applicable, however, to sufficiently large diameter pipes only where the
effects of the wall-near zones are negligible.
The second locus has to satisfy the same value of Ie+,which by definition is

(5.7)

For the same concentration and species, le+ is constant. Thus, in going from a pipe
diameter ofD1 to a pipe diameter ofDz, for the same value ofle+, it follows that

(5.8)
This locus does not exclude the effects of the viscous sublayer and the buffer layer.
Hence, the curved lines of the constant, ~B. have to be plotted at sufficiently close
intervals so that the locus of the constant, ~B. may be interpolated in scaling up data
from one pipe diameter to another. This is shown in Fig. 5.3.
From a measured point a locus of the constant, ~B. is drawn by interpolating between
two nearby lines of the constant, ~B. With eq. (5.8) the values for a new diameter can be
predicted. In the plot, for example, the values for two pipes are given: one with a
diameter twice as large as the test pipe, and the second, four times as large.

5.2 The turbulent velocity field

As mentioned above, we know that at drag reducing conditions an elastic layer forms. It
is one of the goals to understand this phenomenon and the related drag reducing
mechanism. It is hypothesized that the reason for the formation of this layer is an
interaction of the flow field with the polymer molecules which are deformed under these
conditions. Thus, the structure of the flow field changes also by a feedback process,
which ends in a statistically stationary state. In other words, one needs to know the
structure of turbulence in the fully developed case in order to understand the process of
the effect itself. In this respect it is preferable to know the specific structure of turbulence
at onset condition and at maximum drag reducing condition.
DRAG REDUCTION IN POLYMER SOLUTIONS 109

5.2.1 Turbulent structures in statistical decomposition

In a strict statistical sense the first four moments of the velocity fluctuations were
measured and discussed by several authors, whereas to our knowledge an analysis of
pressure fluctuation data does not exist. Moreover, in non-Newtonian flows the pressure
fluctuations are hardly measurable due to additional normal stresses in the fluid.

5.2.1.1 Turbulence intensity

The rms of the velocity fluctuations of one component of the velocity fluctuations, known
as turbulence intensity, and defined by eq. (4.75), was measured for a variety of drag
reducing flows. It became evident that only measurements by LDA were reliable since the
velocity field was not disturbed by the probe and, thus, no effect of a contamination of
the probe occurred.

Wasser !OOppm Pr2360


o Re 10000 Rep ~BOO, DR ~9X
6 !5000 .A. 7600, 6DX
'V 20200 't' !0800. 65X
0 25!00 1~900, 67X

Figure 5.4 Streamwise velocity fluctuations in a pipe flow normalized by Ut in case of


maximum drag reduction, from Gampert & Yong (1990)

The differences found compared with the turbulence intensity in Newtonian flows
occurred in the elastic layer only. The presentation of the results is mainly given in
viscous units since the most characteristic changes occur in the near-wall buffer zone. In
this layer the turbulence intensity in streamwise direction showed a shift of the maximum
fluctuations towards higher wall distances for all measurements, and the peak itself was
broader, see Fig 5.4. Since the location of this peak is equivalent with the turbulence
110 CHAPTERV

production region, one can conclude: In the drag reducing case there exists a shift of the
production towards larger wall distances and the production region is larger.
Additionally, many authors found that the turbulence intensity in streamwise direction is
higher than that in Newtonian flows. However, these data are normally taken under
conditions at which the elastic layer was not fully developed. Gampert & Yong (1990)
presented data of a Polyacrylamide-Copolymer solution in water taken under maximum
drag reducing conditions. These results are shown in Fig. 5.4. The streamwise
fluctuations are lower than for the solvent. They are not universal and strongly depend on
the Reynolds number and increase with ReQ.

Wasser !OOppm Pr2360


" Re =10000 Re = 10700: OR= 57X
0.08 0 14900 + 16000; 70X
.J u'2
UQ o.o6

0
0.04
"
0.02

0. 00
' ' . : ..

:: . . . . . . . ..

l_.JLJ..-'--l....LI._ _---'----'---'---'---'-'.....L.!--'----'--__J'----.J
4 !0 y+ 100 400

Figure 5.5 Streamwise velocity fluctuations in a pipe flow normalized by UQ in case of


maximum drag reduction, from Gampert & Yong (1990)

Since pairs of ReQ for water and polymer solutions correspond to the same flow rate, the
absolute fluctuation values for water and polymer solutions can be compared directly
when the intensity is scaled with UQ. Fig. 5.5 shows that the intensity is drastically
reduced by polymers when UQ is held constant and is scaling fairly universally for a
given polymer-solvent system. These results are consistent with the results of Reischman
& Tiederman (1975).
The velocity fluctuation intensities normal to the wall show all reduced levels, e.g. Yong
(1990) or Luchik & Tiederman (1988). See Fig. 5.6.
The data plotted in Fig 5.6, scaled with ii"Q. show the same behaviour, but the scaling
with u-r is still better.
DRAG REDUCTION IN POLYMER SOLUTIONS 111

Wasser 100ppm Pr2360


6 Re =10000 Re: 10700: DR: 57 X
0 14900 + 16000: 70X
!.2
M
u,
0.8

0
6
0
6


0
6




.... ,..... 0
6 0

0.4
t t
0. 0 L-.J--L...L..l...L.l.----L-..L._--'-....L.._.L..L...L.L..l.-_ _.l.._--'.__j
4 10 100 400

Figure 5.6 Normal velocity fluctuations in a pipe flow normalized by Ut in case of


maximum drag reduction from, Yong (1990)

5.2.1.2 Reynolds stresses

The Reynolds stress tensor (eq. 4.34) term, u'v', is useful as a measure of the
momentum transport in a turbulent flow, see Fig. 5.7. The data for a water flow increase
with wall distance and reach values between 0.4 and 0.5 in a large range of y+. The data
for polymer solutions show that the drag reduction by polymers does not only result from
diminished fluctuations but, in addition, in a decorrelation of the u'- and the v'
fluctuations, see Fig 6.10.
For turbulent flows of a Newtonian fluid the influence of the viscous forces decreases
with wall distance and can be neglected at wall distances of about 80 in viscous units or
about 0.15 in distances scaled with R (This is a value for rough calculations since the
exact values are Re dependent.) For y+ larger than this value the total shear stresses for a
Newtonian fluid are given by the Reynolds stresses. However, measurements of the
Reynolds stresses of dilute polymer solutions exhibit, assuming a constant viscosity a
deficit in the stress balance (see also Willmarth et al. 1987). The stress balance becomes

u'v' y dii+
--=1-----G
u~ R dy+ (5.9)
112 CHAPTERV

Wasser 100ppm Pr2360


0.6 A Re: 10000 Re : 10700; DR : 57X
0 14900 16000; 70%

- U'V' 0.5
Jl?;.{Vii
u

0.3 A

. ... ..
0
0
0.2

:
0
t
0.1
: : *t A
A
tt
0

0.0
4
li1 10 !00 400
y+

Figure 5.7 Reynolds stress profile at maximum drag reduction from, Gampert & Yong
(1990)

In eq. (5.9) G stands for the Reynolds stress deficit or the elastic stresses, as it was
called by Schiimmer & Thielen (1"981). This deficit occurs in flows of drag reducing
polymer solutions whenever the shear viscosity is used for calculating y+, Fig. 5.8.

0
Ill
ci
0
"
ci
0

G(y+)
"'ci
0
N
ci
0
ci
0
0
ci
0

0
I 0.0 10.0 20.0 30.0 40.0 50.0

y+

Figure 5.8 Reynods stress deficit, G, in a pipe flow with polymer additives at maximum
drag reduction from, Schiimmer & Thielen (1981)
DRAG REDUCTION IN POLYMER SOLUTIONS 113

0
0
Ill

0
C!
"
0
0

"'0
, (y+) 0

"'0
0

0
0
0
0.0 10.0 20.0 .30.0 40.0 50.0

y+
Figure 5.9 Effective viscosity, ll, as a function of the dimensionless wall distance from a
pipe flow at maximum drag reduction from, Gyr & Bewersdorff (1990)

The origin of the Reynolds stress deficit, however, is unknown since eq. (5.9) does not
give any information about it. By calling G an elastic stress, one implies that the deficit is
due to elastic properties resulting from rheological effects. However, this is not
necessarily the reason. An alternative would be that the structures of the turbulent flow
are altered so that they can be dominated by viscosity up to higher distances from the
wall. The quantity G, for example, could be interpreted by a wall distance dependent,
effective viscosity, ll

(5.10)

where Veff is chosen so that eq. (5.9) can be written as

u'v' y dii+
--=1---TJ-
u~ R dy+ (5.11)

This is shown in Fig. 5.9.

5.2.1.3 Joint probability density function and higher moments

For a discussion of the momentum transport, as well as theit involved structures, the
joint probability density functions and their plots in the u'-v' planes are, in our opinion,
among the most appropriate ones.
The plot of a joint probability density function , (JPDF), shows the probability of finding
the velocity fluctuation vector, y', pointing into an area element of the u'-v' plane at a
given time, where the probability is shown by contour lines of the JPDF, see Fig 5.10.
114 CHAPTERV

v' v'
y+= const.

P(v')

u'

P(u')

Figure 5.10 A joint probability density plot taken at a constant wall distance where the
probability function is already centrally symmetric. The explanation of how to
read such a plot is provided in the text

The largest contour line is given by an arbitrarily chosen small probability value. In the
interior the contour lines are given by lines of constant probability increased by an
increment, tl. Normally, tl is held constant. The result is that the JPDF shows up as
coaxial ellipses, with their center at the origin.
These ellipses, however, appear at a certain wall distance only where they also become
centrally symmetric. This is not the case close to the wall, where the contour lines are no
longer ellipses but still look like one-sided wrenched ellipses.
The ratio of the two principal axes is a measure of the deviation from isotropy of u' and
v' at the measuring location. If the ellipses degenerate into coaxial circles, the turbulence
becomes in a two-dimensional sense isotropic at that location. This can be found at larger
distances from the wall, y+,., 2-300.
DRAG REDUCTION IN POLYMER SOLUTIONS 115

The axes of these ellipses are turned in respect to the u'-v' axes by an angle a. The
distributions of u'v' is no longer symmetrice in respect to the coordinate axes u'-v'. In
turbulent flows the second and fourth quadranta are dominant. We will return to an
interpretation of these values of the four quadrants when discussing the structural
behaviour. The angle, a, provides the direction in which a fluid particle moves in the
mean, because the mean probability of the orientation of :i is in the axi-symmetric case
exactly given by a.
Since the frequency with which the vector position is recorded as an event to define the
probability density is a constant, an integration in stripes parallel to the coordinate axis
directly gives the time probability density, P(u') or P(v'), of finding a fluctuation u' or v'
of the size defined by the distance of the stripes.

Wasser IOOppm Pr2360


t:c. Re =10000 Ae: 10700: DR: 57X
!.0
Sx
0 14900 + 16000: 70X

2L-~-L~~----~--~-L-L~LL~-----J---L~

4 10 100 400

Figure 5.11 Skewness, Sx. and flatness, Fx, in flow direction. From Yong, (1990)
116 CHAPTERV

y' = !0 yf ::: !0
...::
0
~

~
'~--'=
~
_-::::_

-
-=- .) ~
~
4 ~
~

-3

y' = !5 r' ~ !5

~~ -~"
+;

--
~ '<;

~ ~----=-v
-3

~ r' =15 y' : 15

~~~:::.- ~
~~
-uoV --~
~..-.,..,~

-~
=____,.~"->---;
-i?-'.3?

-3

!~
y' = 35
/~~~~~
--~
':.<-;:?/

-3 I
y+ = SQ
(~ ~ ~50
~ "'-
r:;<-;;.~ ~~
~~~~
r- ~~
-3
y1 = 70
~=70
~ ~--c
'4-~- ~~ ~ ...

-~ S2)) ~~ ~~

-3
3
y' = !DO
~!00
~ - ,__
~~
~~~ ~'?__/

r- ~~ J ~

-3 I

-ur
-5 -5 U'

Figure 5.12 JPDF in the u'-v' plane for flows at Re= 15000 at different wall distances,
with a minimum contour line of 0.01 and increments of 0.025. For water to
the left, for 100 ppm Praestol 2360 from Stockhausen CO to the right. From,
Gampert & Yong (1990)
DRAG REDUCTION IN POLYMER SOLUTIONS 117

The higher moments of the velocity fluctuations, u' and v', can be calculated by using
their time series or by considering them as the moments of P(u') and P(v'). From this it
is evident that, in case of contour lines of coaxial elliptic form, the P's become
symmetric, and, therefore, all odd moments have to disappear, especially the skewness.
The values of the skewness, eq.(4.77), due to eq. (4.76 & 4.75), and of the flatness or
kurtosis, eq. (4.78), in flow direction of the data of Yong (1990), are given in Fig. 5.11.
They are a measure for the symmetry and peakedness of the probability density
functions.
With this introduction into the powerful tool of presenting the u' and v' data, the data of
Gampert & Yong (1990) are used again to show a complete set of data analyses based on
the same measurements, Fig. 5.12.
The significant differences between the JPDF of the solution and the solvent are:
The area surrounded by the minimum contour line for a polymer solution is
smaller and shows that the fluctuations are more directionalwise organized. In
other words, one would expect structures which are more self-similar.
The turbulence is more anisotropic, the u' 's are predominant, and the v' is
significantly smaller for polymer solutions. This finding, together with the result
that also the cross-correlation of the two components of the velocity fluctuation, u'
and v', is strongly reduced, confirms that the structure of the turbulent flow has
drastically changed. The turbulent structure in dilute polymer solutions is
characterized by the dominant fluctuations in mean flow direction.
The angle, a, is larger, which means the main axis of the ellipsis becomes more
parallel with the flow direction. This means that the structures have a relative
motion preferred which is practically parallel to the wall in mean flow direction.
Therefore the fluctuations are transporting less momentum towards the wall. For
structures of the same intensity this is equivalent to a drag reduction.
The skewness as well as the kurtosis are larger. They show that the intermittent
character of the two flows is significantly different. The flatness shows that the
intermittency has increased. The skewness shows that the intensities of the
turbulent "spots" become stronger in the dilute polymer solutions.

5.2.1.4 The kinetic energy distribution, production and dissipation

As formulated in chapter 4.2.2, turbulent kinetic energy is usually characterized by the


power spectrum, E(k), eq. (4.88), as a function of wave numbers, the length scales of
the structures, or in form of its temporal counterpart: the power spectral density, S(ro),
given by eq. (4.91), or the Fourier transform of the autocorrelation, R('t), eq. (4.60). A
118 CHAPTER V

typical normalized Sx(ro) in flow direction, made dimensionless with u't2, is shown in
Fig. 5.13.
It is characteristic for a polymer solution that the energy content in the vortices of
different sizes is shifted from the small to the larger eddies compared with the solvent.
These results were interpreted as a damping of the small eddies and therefore a damping
of the dissipation due to the polymers, Berner & Scrivener (1980).

S (co) 1 H10 PR 2850 H10 PR 2850


1H70 12120
__..._ --"""9---
---- ---o---
__.,_ --~--

c =100ppm

10~~---4----~----+---~
1 10 10 2 103 104 1 10

Figure 5.13 Energy spectra for polymer flows of two different concentrations of PR2850
at three different wall distances; y+<y+ max.y+,y+max andy+= R+, withy+max being the
location at which the turbulent energy becomes a maximum. From, Delgado (1987)

An equivalent statement can be given in terms of microscales, especially in the Taylor


time micro-scale, eq. (4.68), defined by the autocorrelation function. This microscale
increases for polymer solutions, showing that the smallest vortices become larger
compared with Newtonian flows.
The energy available from the pressure gradient is balanced by direct viscous dissipation,
Dv. and the caused turbulent production, P, eq. (4.42). This part will finally also be
dissipated by viscosity. In pipe flows these two terms can be derived from eq. (4.39) as

(au)
Dv =V ()y
2

(5.12)
DRAG REDUCTION IN POLYMER SOLUTIONS 119

P =Pr = -u'v' -(au)(}y


(5.13)

Virk (1975) estimated these values and compared them with Laufer's data (1954) for the
Newtonian case. These results are presented in Fig. 5.14

Pr Dv

MOxHY',um Drag Reduction


0 0 Newton:on

or 0.5

50 100 150 y+

Figure 5.14 Estimated direct viscous dissipation and turbulent energy production profiles
in the wall-near region during Newtonian and maximum drag reduction flows
at Re+= 1000. From, Virk (1975)

The plane of peak turbulent energy production at which flv = 1 shifts from y+ "' 12 to
28. Due to the relation (4.45) the total turbulent energy production must equal the total
dissipation. From Fig 5.14 it becomes evident that the Newtonian dissipation is restricted
to occur in the wall-nearest layer y+,30-40, where for polymer solutions the dissipation
is significantly altered in the sense that viscous dissipation occurs up to wall distances of
about y+, 100-150, a result compatible with the found effective viscosity defined by eq.
(5.10).
120 CHAPTERV

5.2.2 Turbulent structures viewed as "coherent structures"

5.2.2.1 The structures detected by velocity measurements

The differences found in chapter 5.2.1 can be interpreted by arguments based on coherent
structures. As mentioned in chapter 4.2.3, the terminology "coherent structure" can be
misinterpreted, and it is, therefore, important to specify which kind of definition is used
in a particular case.
Here, as a first step, the most simple classification is used, the quadrant splitting. This
description is practically identical with the representation used for the JPDF. It is the
representation of the velocity fluctuations in the (u' ,v') coordinate system. The JPDF, at
least for the Newtonian case, is asymmetric with respect to the coordinate axis. The
probability of finding y' in the second or the fourth quadrant is much higher than for the
two other quadrants. Can this fact be interpreted by the existence of particular structures?
Already Prandtl, by defining his mixing length concept, made use of a structural idea. He
interpreted the main momentum transport as due to an exchange of outer- and near-wall
vortices like fluid blobs. The quadrant splitting makes use of a definition of the structures
very similar to the one by Prandtl. The velocity fluctuation vector, .Ji.(t), is given as a time
series, and this series can be subdivided into events. Their boundaries are defined by the
crossing of .!.L from one quadrant to another, or in other words, when u'v' changes its
sign.
The new structural interpretation by coherent structures is based on a pattern recognition
technique by which the time series of the velocity fluctuations have been analyzed. By
quadrant splitting, together with conditional sampling techniques, the wall-near turbulent
structures could in the mean be interpreted to belong to three different classes. Very close
to the wall there have been found pairs of counter rotating streamwise vortices which
interact with the outer flow by a bursting mechanism due to an instability process of the
wall-nearest flow. This bursting mechanism is characterized by two distinct events, the
sweeps and ejections. In the new terminology the momentum transport is mainly due to
these two events, i.e., Blackwelder & Eckelmann (1979). Sweeps and ejections have a
fairly well defined direction, a, and one can conjecture that they are eventually controlled
by some larger structures or are even part of a larger structure. The interaction with the
outer flow, however, needs a structural counterpart in this region. As such, the so-called
big structures or eddies were postulated. A model of how this interaction can be imagined
can be found in Laufer (1975) and has to a certain extent been verified by the observation
of Chen & Blackwelder (1978), who, by using temperature as a tracer, found that the
bursting phenomenon was observable as fronts of shear layers throughout the bulk flow.
The difference between this description and the previous ones is that the new one is more
DRAG REDUCTION IN POLYMER SOLUTIONS 121

highly deterministic. But it has to be warned, all these structures are defined by their
mean behaviour.
The JPDF also shows that the ~a between sweep and ejection is 1800, or geometrically
interpreted, the ejections seem to be like a reflection of a sweep reflected at the wall. The
additives mainly change the a and, therefore, the angle of "reflection." In polymer
solutions for sweeps, a increasingly approaches the value of 1800. In the case of a=
180 the momentum transport for coaxial elliptical JPDF becomes zero. Therefore the
question arises, by which mechanism is the momentum transport towards the wall
maintained?
In the Newtonian case the momentum transport to the wall differs with wall distance. The
dominant momentum transfer is due to Reynolds stresses, and only in the region very
close to the wall is it due to viscous stresses. This region is called the viscous sub layer,
with u',v',O.
In the drag reducing case the fluid still exhibits fluctuations, though they are anisotropic
v' << u'. The reasons for this anisotropy can be of different origin. The common
interpretation is that, since u' still remains large, one can assume that the dominant
structures are large structures and the fluctuations are produced by their relative velocity.
Further away from the wall the JPDF becomes more isotropic. This is the case for
Newtonian fluids as well as for polymer solutions, but not for flows of maximum drag
reduction, Gampert & Yong (1990).
A very important piece of information is the intensity of an event. It can be calculated by
integrating u'v' over the total event. This intensity, IE, is given by

IE= p Ju'v'dt
TE (5.14)

Schmid (1985) found that the mean intensity ascertained from the mean over all IE, which
was found by eq. (5.14), differs most in the ejection type events. Up to a wall distance of
about y+,20 the turbulence is higher for water than for polymer solution flows.
However, in the buffer zone, IE can be increased by up to 40% in flows of polymer
solutions. Schmid's data were taken at low drag reduction. At higher drag reduction
Schmidt (1989) found the same behaviour for the ejections; for the sweeps in polymer
solutions, however, he found much higher values in the buffer zone, whereas in the wall-
nearest layer no significant difference was detected. In the core region (logarithmic part of
the profile) no differences between the two flows were observed.
The intensity itself, however, is not sufficient for deciding how the momentum is
transported. It is also necessary to know the duration of the event. Schmidt (1989)
showed that in all cases the duration of the events for polymer solutions is longer,
122 CHAPTERV

increases with drag reduction and occurs from the closest distance of the wall to the
center of the flow. This is a strong argument against models which assume an overall
damping of the turbulence by additives. Together with the results from power spectra or
the autocorrelation function this result can now also be interpreted in terms of structures.
The momentum transport in polymer solutions is due to structures of larger length scales
with longer duration, Berman & Cooper (1972).
Besides these rather crudely defined structures, the only well established coherent
structures are the longitudinal counterrotating vortices close to the wall, and they too are
structures which were detected only in the time mean. As Robinson (1990 & 1991)
showed even very close to the wall there exist a wide variety of specific structures. Yet,
in spite of these critical remarks, it is widely assumed that these longitudinal vortices are
the main elements responsible for the instability processes, which end up in the
turbulence production. The streaky pattern of low and fast fluid separated by these
vortices has for Newtonian flows a wavelength across the flow ofXz+"" 80-100. The
values for polymer solutions vary but are reported to be about 20% higher. The reason
for these differences is very complex, and, therefore, some simulation calculations have
been performed to check how specific rheologies produce larger longitudinal vortices.
The most simple approach is to use an effective viscosity, TJ, eq. (5.10), Gyr &
Bewersdorff (1990). A more sophisticated constitutive equation was introduced by
Lyons et al. (1988). Both rheologies showed an increase in the vortical sizes.
All other modifications of "coherent structures" cannot be predicted or explained.
Nevertheless, information on the altering of the structures is extremely valuable, and one
can hope that in the near future more investigations with other rheologies will be made.
These investigation, however, will hopefully study more than only the behaviour of the
structures in a quadrant decomposition.

5.2.2.2 Topological aspects of drag reduction

The concept of "coherent structure" received a new meaning in the studies of the
topological aspects of the turbulent flow field. For homogeneous isotropic turbulence,
Fung et al. (1991) found a set of objective criteria, in the kinematic case based on the
local strain rate, vorticity and pressure, which describe typical regions in which the
streamline circulates, converges or diverges, and forms streams of high velocity flow.
Although restricted to the kinematic approach, Fung et al.'s (1990) classification is a very
useful one, since the characteristic flow zones exist even if the flow field is studied in its
full dynamical form. In a generalization these zones can be thought to be the basic
elements or the "coherent structures" in the core flow. The interesting part of such a
description is that they give local micro flow fields, and these are the flow fields which
DRAG REDUCTION IN POLYMER SOLUTIONS 123

interact with the additives, since the most active flow regions around the critical points of
the flow field are of the scale of the elongated polymers. If the additives are deformed in
these fields and change the rheology, one can estimate the local influence. However, even
with such information we are still far from a complete description because the feed-back
of the local changes of the structure on the flow field is still missing. Nevertheless, this
model of the turbulent flow field allows to check some of the hypotheses on how drag
reduction occurs.
The influence of additives on the topology of the turbulent flow field has not been
investigated yet. What has been investigated is the interaction of polymers with well
defined flow fields mainly used to investigate the rheological properties of the fluid. It,
therefore, seems worthwhile to give the criteria used by Fung et al. so that the results of
these isolated experiments can be interpreted.
They used three quantities to classify the zones: the second invariant of du'Jihj, 11, the
pressure, p, and the kinetic energy, 1!!.'12.
The second invariant can be represented in terms of the strain rate, dij, and the vorticity,
Wk

(5.15)

In other words, it is the divergence of p or the source of the pressure force field which is
needed to hold the kinematically calculated flow field in equilibrium. This is the case,
however, only by assuming that v= const.without any non-Newtonian effects

(5.16)

The pressure can, therefore, be obtained from the flow field by using the Poisson
equation (4.37). 11 is, therefore, equivalent to specifying higher derivatives of the
pressure, p. A maximum in p occurs where 11 is negative. However, to ensure that the
zone occupies a significant volume, a second criterion has to be formulated, based on the
pressure determined by the average volume or 11 over a certain region around the point of
interest.
Following Wray & Hunt (1990), four flow classes are distinguished, Fig. 5.15, using
eq. (5.15) for the evaluation of 11, eq. (4.37) to determine p by the velocity field, and lui
as the square root of the local kinetic energy. These criteria are mutually exclusive, and,
in fact, about half of the flow field remains unclassified.
124 CHAPTERV

Eddies (E) Shear (SH) Streaming (S) Convergence (C)

TI< -1/2 Tirms TI< -1/2 TI rms ITII< 1/2 TI rms TI> TI rms
p < -1/2 prms -1/2 prms< p< Prms lui> Urms P > Prms

Figure 5.15 The classification of the characteristic turbulent zones defined by rms values
taken as global average in homogeneous, isotropic flows and the average
over a plane of constant shear rate in shear flow. From Fung et al. (1991)

For a low source term (negative values of ll) and low pressure the streamlines tum into a
focal node. This type of flow is called eddies, (E), and at moderate pressure the
streamlines become more parallel, and one calls these features shear layers, (SH). At
moderate values of the source, in pressure forces but high kinetic energy, the flow
becomes jet-like or "streaming", (S). If the source becomes strong and the pressure high,
the streamlines converge towards, for example, a stagnation point. The flow fields are
called convergence fields, (C).
This classification of the flow field is very similar to a topological classification due to
critical points, see Perry & Chong (1987).
To classify the strain in the different' zones, the third invariant of the strain tensor, dij.
defined as

(5.17)

can be used. When S>O, a material sphere will elongate, and when S<O, it will be
flattened.
For a more dynamical topological interpretation of the turbulent flow field, Moffatt's
model (1990) is a good example. In this model turbulence is regarded as a weakly
interacting sea of vortons. As Moffatt (1993) pointed out, spiral structures are the natural
candidates for the role of the "generic structures" of turbulent flow because they are
eventually outcomes of a Kelvin-Helmholtz instability, an all-pervasive phenomenon
associated with nearly all shear flows at very high Reynolds numbers. Such structures
were proposed by Lundgren (1982) in a model of the fine structure of turbulence, in
which axial stretching of rolled-up spiral vortices plays an essential role.
This new view seems to be very important for the understanding of drag reduction. Thus,
we would like to make a short excursus on some ideas used for theoretical explanations
of the effect. The common explanation was that additives act as drag reducers when their
aspect ratio is large enough. Hinch & Elata (1979) concluded that for surfactants and
fibres this is normally the case; polymer molecules, however, have to be stretched before
DRAG REDUCTION IN POLYMER SOLUTIONS 125

they can act. For the stretching of the molecules, high strain fields are needed. It was,
therefore, conjectured that the polymers become stretched when, in the turbulent flow
field, zones of high strain are present, an idea which is in agreement with the onset
behaviour of the effect. Also in agreement with this model is that the effect starts in the
buffer layer, where the strongest strain fields are present in the form of ejection-like
events. Once the polymers are elongated by such a structure, they could act like the other
drag reducing additives. However, the polymer molecules need time to be stretched and
they recoil again due to Brownian motion of the different segments of the molecule.
Based on the knowledge of the two characteristic times, i.e., the stretching time and the
relaxation spectra, the amount of stretched material can be estimated if the frequency at
which the ejections occur is known. The exposure time should be sufficiently long and
the bursting period, the time between bursting events, short in comparison to the
relaxation time, eq. (2.13), of the polymers. This is not the case, and, therefore, many
explanations were based on an inhibition or damping of the stretching mechanism itself.
Such a damping could already be drag reducing since the ejections are of helical flow type
and are therefore, sources of high vortex stretching, the main mechanism for the
enstrophy generation, eq. (4.97), and the result of high dissipation, eq. (4.54). Reducing
the vortex stretching is drag reducing, but this mechanism is not restricted to the ejection
type of flow. Therefore this explanation is too simple and cannot, for example, explain
why an onset behaviour is necessary for the effect. Thus, the mechanism has to be more
complex, while more stretched material must be present.
The topological description shows that the flow field is composed of many more zones,
which are at least capable of maintaining the stretched form and transporting elongated
material into new strain fields before they can relax. In other words, the flow field is
much more complex and not only composed of burst events and is, therefore, very
capable of maintaining a drag reducing state.
An indication of a more complex molecule flow interaction is that polymers which were
presheared in a laminar Couette flow exhibit a higher resistance or elongational viscosity
in an elongational field, a result reported by James et al. (1987), Wunderlich & James
(1987) and Vissmann & Bewersdorff (1990). This is in agreement with Abernathy's
result (private communication), who could prove that polymers decoil very efficiently in
shear or even stronger in helical shear flows. Therefore, the eddy (E)- shear (SH)- and
streaming (S) zones can all enhance the elongation mechanism as well as the maintenance
of already stretched molecules.
In a first attempt to use such ideas, Den Toonder et al. (1993) could show in a numerical
simulation of a turbulent pipe flow that if the polymers just acted by their elongational
viscosity and damped the zones of high strain, almost no drag reduction occurred, thus
giving evidence that the effect is due to a rather complicated interaction mechanism.
126 CHAPTERV

5.2.2.3 Visualization of structures

Three categories of visualization related to drag reduction by polymers are available. (1)
Most of the visualizations are indirect information as they deal with optical investigations
of rheological properties of the solution or they investigate specific laminar flows like
flows through nozzles and four roll mills. (2) The investigations of separations are also
informative for the turbulent flow behaviour of such fluids, but are still an indirect study
with respect to drag reduction. (3) The third category deals with the visual studies of
typical structures in a drag reducing turbulent flow due to the polymers. Only very few
publications of this kind are known. Achia & Thompson (1974,1977) showed by
holographic interferometry that the mean spacing of the streaks oriented in flow direction
was increased and accompanied by a decrease in burst frequency. The ejection type
structure showed a definite suppression of the elongational flow field. A similar result
was found by Donohue et al. (1972) by using dye injections. A typical visualization
picture was given by Tiederman (1974). Using fine sand as a wall-near tracer, Gyr &
MUller (1975) showed that besides the longitudinal wave pattern in drag reducing flows,
a transverse wave pattern is also stabilized due to polymers, and the instability produced
by a tripping wire remains two-dimensional much longer in the transverse direction.
However, to our knowledge there exists no additional visualizations which could be used
for a quantitative determination of the main features of drag reducing flows.

5.3 Heterogeneous drag reduction

For a scientifically clean experiment it is necessary to know if the polymer molecules are
monodisperse and isolated (dilute) or not. In the case that they form agglomerates, one
never knows the dimensions of the additives, how they are deformed, and what is their
number concentration, i.e., the number of molecules per unit volume. However, for
practical applications the premixing of the solution necessary to achieve a dilute state is
very tedious, and therefore the polymers are injected in higher concentrations, assuming
that this fluid will be dissolved so rapidly that the dilute state is realized in the test section.
However, concentrated polymer solutions form entanglements between the molecules.
According to Barnes et al. (1989), "entanglements are not merely the overlapping of
adjacent molecules since overlapping occurs at quite low molecular weight. Rather
entanglements are strong couplings, whose details are largely unresolved, but which act
in a weak manner like chemical cross-links between the molecules." These entanglements
cause a strong viscoelasticity.
DRAG REDUCTION IN POLYMER SOLUTIONS 127

Nevertheless, the injecting technique was also widely used for scientific investigations.
The polymer solution was injected through slots into the wall region of the flow or
through a nozzle into the core region of the flow. Most publications are based on
experiments of this kind. Even the data of the papers used as standard values are often
raised by these techniques. It, therefore, becomes understandable that they were criticized
as being results found for solutions which did not contain isolated molecules, e.g.
Dunlop & Cox (1977), but rather for solutions of polymer aggregates. This is the reason
why in chapter 5.2 all papers based on an injection technique were avoided. In our view,
experiments of this kind are experiments with heterogeneous solutions, a result
confirmed by the work of Hinch & Elata (1979). For a time dependent network system
they found that the onset depends on polymer concentration and that the time constant of
the network is the ratio between shear modulus and viscosity of the concentrated solution
of entangled molecules with which the network is created. In the homogeneous dilute
solution, below the saturation concentration, onset is independent of concentration and its
time constant is typically of the order of the largest molecular relaxation time.
One very important result was found very early by this technique. Wells & Spangler
(1967) found that polymer solution injected at the wall of the pipe flow reduced the wall
shear stress immediately, whereas when the same polymer solution was injected into the
turbulent core region, no effect was observed until the polymer had diffused to the wall
region. Me Comb & Rabie (1982) claimed that this region is limited by 15<y+<JOO.
The terminology heterogeneous drag reduction is usually used for drag reduction
achieved by injecting a concentrated polymer solution into the core region of the turbulent
pipe or channel flow. The concentration is usually chosen due to the known parameter of
the polymers and the solvent, e.g. a 0.5- 1% polyacrylamide or polyethylenoxide
solution in water.
Depending on the concentration of the injected polymer, the solution remains more or less
a thread or disintegrates into smaller pieces depending on the original concentration of the
injected solution. If c* is the limiting concentration between a dilute and a semi-dilute
solution, the material dissolves into single molecules or into aggregates equivalent to
"super-molecules," or remains a thread depending on c<c*, c"'c*, or c>c*. The
mechanism/s of dissolving is only to a small extent a diffusion process that means a
molecular process. More often the thread splits off or can even be disrupted into pieces
which finally dissolve. A certain classification of this can be found in Bewersdorff
(1989) or in Usui (1990). This mechanical dissolving process is of importance since
those researchers who explain the drag reduction as a direct interaction of the flow with
the thread argue that the effect due to the thread appears only if the thread remains very
compact and that in all other cases dissolved material is the cause. It is the case of a
compact thread which has been debated since, according to one opinion, the drag
128 CHAPTERV

reduction is due to the interaction of the viscoelastic thread with the turbulence of the
Newtonian fluid, while according to the other opinion drag reduction is due to minute
amounts of polymer which are removed from the thread and dispersed into the bulk fluid.
A number of observations have shown that the thread preserves its identity for
downstream distances of many pipe diameters as long as c>c*. Also for this case it has
been proven by Smith & Tiederman (1991) or Gyr et al. (1993) and Hoyer (1994) that
the effect is produced by dissolved material, and therefore this effect is similar to
homogeneous drag reduction by dilute polymer solutions, which shows up in an onset
behaviour of the effect, as was found for homogeneous polymer solutions in the limits of
dilute solutions defined by Hinch & Elata (1979). For a long time this explanation was
rejected as it did not seem probable that the small amounts of material which can possibly
leave the thread could produce the strong effects which are observed for a single
continuous thread. An upper bound for the concentration of dissolved material in the
near-wall zone was given by Bewersdorff (1985) as 1 ppm. Hoyer (1994) found that the
effective concentrations are as low as 0.1-0.2 ppm. Homogeneous drag reduction does
not normally occur at such low concentrations, and the observed reduction presumably is
caused by agglomerates. By a light scattering technique the active additives were
measured, Gyr et al. (1993), with an Ra of the order of 225 nm or about twice the size of
a single molecule which has a radius of gyration of the order of 180 nm. The effective
material seems to be composed of couples of entangled molecules which act as if they
would be "super molecules," or extra long molecules, from which it is known that they
are drag reducing at such low concentrations, Olivier & Bakhtiyarov (1983).
It is clear that the thread may also influence the transport properties of the turbulent flow.
There is no evidence, however, that this interaction is drag reducing. The argument that
drag reduction occurs due to forcing the turbulent structures by the thread was thought to
be the reason. However, in the power spectra neither a forcing frequency nor a resonance
frequency could be found. Experiments with a rubber thread of very low elastic modulus,
Hoyer et al. (1994), confirmed that even such threads are unable to produce drag
reduction. What remains to be explained is the anomality of the velocity profiles.
Many different behaviours of the velocity profiles were found in the past, depending on
the contribution of the different parameters already mentioned. At the end of the viscous
sublayer, which seems to be unchanged in comparison to the Newtonian case, the profile
for very compact threads follows the Newtonian curve of the turbulent core. At y+, 30,
however, the slope of the velocity profile increases to give larger values of u+. The slope
in this transition layer is higher than in the turbulent core, i.e., no parallel shift can be
found, but instead, a curve with slope smaller than in Virk's ultimate profile, Fig. 5.16
(a). For more dilute threads, instead of a transition profile in an elastic layer ending up in
a Newtonian plug flow, one finds a large elastic layer of the size of the cross section with
DRAG REDUCTION IN POLYMER SOLUTIONS 129

a logarithmic profile with a slope increment between the Prandtl-von Karman law and
Virk's ultimate profile, profiles which are equivalent with the one found by McComb &
Rabie (1982), see Fig. 5.16 (b). The slope increases with increasing Re. There seem to
be two reasons for this dependence. At higher Reynolds numbers more material is
dissolved, and the ratio between the length scale of the molecules and the turbulent flow
structures increases with increasing Re. From dilute homogeneous polymer solutions we
know that if this ratio increases, the efficiency of the effect increases too. This statement
is in agreement with the findings of Me Comb & Rabie (1982), who injected a solution at
the centreline, which, because of its fairly low concentration, was thought to immediately
disperse and become quasi dilute after a rather short distance. In this experiment Me
Comb & Rabie found that the increment of the slope in the velocity profiles increases
with the distance from the location of injection. Thus, the dissolving process is of
importance because the concentration of "active" additives increases downstream. What
we know is that the efficiency of the "additives" at drag reduction depends on their aspect
ratio and other molecular parameters: Therefore, "active" stands for the appropriate size.
Hoyer et al. (1992) showed that agglomerates of the right size ("active") are dissolved
most efficiently when the thread is strained and contracted.

40

viscous-
30 sublayer

10

0
1 0 100 1000

Figure 5.16 A schematic representation of the profiles in case of heterogeneous drag


reduction (a) for compact threads (b) for dilute threads and threads which
disperse into pieces just after the nozzle
130 CHAP'IERV

Although these results are in favour of an explanation based on dissolved material,


several researchers still argue that at least part of the effect could be due to the direct
interaction of the thread with the flow, Hoyt & Sellin (1991) or Usui et al. (1993), while
admitting, however, that the direct effect of the thread can be only small. Usui et al.
explained the effect by an elastic storing effect, although they were not able to explain the
Reynolds stress deficit in the near-wall region where this is a maximum.
While it might seem to be only of academic interest, the discussion of this problem is of
basic importance since a direct effect of the thread would be a new kind of a drag
reducing mechanism. Therefore, several attempts were made to investigate the problem,
including a combination of visualization and simultaneous LDA velocity measurements.
The turbulent flow exhibits forces distributed over the thread surface. Due to force
gradients, the thread is moved and deformed. It can be observed that the thread becomes
elongated in the mean; locally, however, the thread sections can relax due to elastic
forces. Therefore, an effect cannot be denied a priori, and one of the authors,
Bewersdorff, is of this opinion. The other, Gyr, believes that, based on the energy
budget, no drag reduction due to a direct interaction is present.
The most interesting result of this type of experiment is that, in case of heterogeneous
drag reduction, the velocity profiles have the character of a Virk profile. However,
although the elastic layer occupies the complete bulk area, the velocity profiles have a
lower slope than Virk's ultimate profile. In other words, it can be speculated that the best
drag reducers are the monodisperse, homogeneously distributed polymer molecules. On
the other hand, since the amount of material producing the reduction is so extremely low,
it can be argued that the efficiency with respect to the cost is by far the best in the case of
a heterogeneous injection.
A heterogeneous drag reduction experiment, on the one hand, does not seem suitable for
basic research. Since heterogeneous drag reduction is the most efficient for applications,
it will also be a theme in future.
This chapter would be incomplete without mentioning a very unusual result. Me Comb &
Rabie (1982) observed that, in the mean, the drag reduction itself, which is measured as
the difference in wall shear stress, oscillates along the pipe if the solution was injected at
the wall, and the oscillation scales with the burst frequency. Since this is a result in time
average, this can only be understood as a triggering of the burst cycle by the disturbance.

5.4 Drag reduction under rough wall conditions

Up to this point only drag reduction in smooth pipes has been discussed. The results
found were mainly interpreted as due to changes in the turbulent structure. Therefore, an
investigation of the turbulent flow of dilute polymer solutions in rough pipes would be a
DRAG REDUCTION IN POLYMER SOLUTIONS 131

good test for several hypotheses regarding how far specific alterations of the turbulent
structures are responsible for the effect. For example, if the change in the wall-nearest
structures, i.e., in the longitudinal vortices, is the main reason for the drag reducing
effect, as has been claimed by several authors, then drag reduction should disappear or at
least be significantly reduced in rough pipe flows.
From the point of application, a drastically reduced effect would be a severe restriction
since most pipes used under conditions where drag reduction occurs are normally rough.
(ut at onset"' 25 mm/s means that in water y+=1 corresponds toy of 410-2 mm; if rough
is defined as y+>5, then a roughness of 0.2 mm means as a rough wall.) This estimate
stands for sand roughness. Besides this kind of roughness, d-type roughness can also
occur. For d-type roughness, the effect of rough surface on the turbulent flow is
dependent on the channel width and not only on the dimension of the roughness element.
In that case, the roughness length becomes a fixed fraction of the channel width, d. A
typical d-type roughness is, for example, a rough surface composed of similar
rectangular bars laid transversely to the direction of flow with spacing equal to the width
of the bars. For the classification of the roughness types, see Perry et al. (1969). Since
the configuration of the roughness elements is also important for the actual drag, we have
to distinguish even for sand roughnesses between uniform and non-uniform
roughnesses.
Despite these very important arguments pro studying drag reduction in rough pipes, such
studies were very seldom done. Lindgren & Hoot (1968) first detected drag reduction for
a 60 ppm Polyox solution in a rectangular channel with rough walls. In general, drag
reduction in rough pipes exhibits a maximum at a certain Reynolds number. McNally
(1968) detected that maximum drag reduction always occurred at the same friction
velocity, regardless of the polymer concentration, in agreement with the value k+, 50
found by Virk (1971), where k stands for the heights of the roughness elements. The
decrease in drag reduction at high Reynolds numbers beyond the maximum was related to
mechanical degradation of the polymers due to the influence of the roughness, Debrule &
Sabersky (1974). By contrast, Spangler (1969) and Virk (1971) found hydrodynamic
effects to be responsible for this effect. Spangler (1969) mounted a smooth pipe
downstream from a rough pipe, and measurements in the smooth pipe demonstrated that
no degradation had occurred. In Virk's (1971) experiments the pressure drop at a
Reynolds number above the maximum drag reduction remained nearly constant at a
constant flow-rate in a closed loop after several runs and did not increase as one would
have expected in case of mechanical degradation. A systematic study of the friction
behaviour of dilute homogeneous polymer solutions was done by Virk (1971).
Virk's result clearly shows that drag reduction also occurs in rough pipes. He
investigated the flow at a relative sand-roughnesses of R!k = 15, 23, 35 in pipes of an
132 CHAPlERV

inner radius, R, of 43 mm. Thus, the heights, k, of the roughness elements were k= 2.9,
1.9, 1.2 mm, which means the pipes were fairly rough. The onset of drag reduction in
the rough pipe occurred for the same polymer at the same wall shear stress as in the
smooth pipe; the onset wall shear stress was essentially independent of polymer
concentration and varied inversely as the square of RH and was unaffected by the regime
during which onset occurred, whether hydraulically smooth, transitional or fully rough.
Following onset, a regime was observed wherein the fractional slip, Sp,

(5.18)

a function similar to ~B. which was obtained with a given polymer solution in a rough
pipe, was the same as in the smooth pipe despite marked differences in the respective
rough and smooth friction factors. Virk called this the "effectively smooth" regime since
it prevailed for values of non-dimensional roughness, k+* <k+<k+ es. from onset, k+*, to
an upper limit given by k+es"' 50 in all of his experiments.
For k+>k+es the observed Spin the rough pipes was always smaller than that for smooth
pipes and was a function of the relative roughness as well as of flow and polymeric
parameters, but Virk did not find a general trend regarding the nature of drag reduction,
as k+ is increased beyond k+es
The maximum drag reduction possible in the rough pipes was limited by an asymptote
which was independent of polymeric parameters. All the rough pipes identically obeyed
the smooth pipe friction factor relation for k+<12 and had the same maximum drag
reduction asymptote. The onset of the roughness effects at k+,J2 indicated that the
maximum viscous sublayer thickness attained during drag reduction was approximately
2.5 times the Newtonian value.
For a non-uniform roughness, such as is found in commercial pipes, the friction-factor-
Reynolds-number behaviour of a Newtonian fluid can be matched with the non-uniform
roughness data to give an apparent roughness height, ka. Roughness heights of different
pipe materials in handbooks or used in experimental studies with commercial pipes are
measured this way.
When the friction behaviour of drag reducing polymer solutions is studied over a
sufficient range of Reynolds numbers, at a certain Reynolds number a maximum in drag
reduction is found. In the following, evidence is given that this maximum in dra~
reduction and the lower drag reduction at higher Reynolds numbers, respectively,
correspond to transition from smooth to rough friction behaviour in the friction behaviour
of polymer solutions. By an analysis of available literature data, Bewersdorff & Berman
DRAG REDUCTION IN POLYMER SOLUTIONS 133

(1987) found that the viscous length scale, v/u't, at this maximum drag reduction was
always the same, independent of the polymer-solvent system and other polymer
parameters. If ka is not changed by the polymers, then ka+ ranges from 1 to 2 for the
polymer solutions at the transition point. That the polymer solutions, their concentration,
and their molecular weight distributions differed in the various experimental studies
suggests that a mechanism involving a polymer time scale is not relevant. Thus,
degradation of the polymers is an unlikely explanation.
For Newtonian flow the transition occurs at ka+, 3. The reduced critical value of
roughness corresponding to ka+"' 1-2 would imply that the heights of the roughness
elements are much smaller than the thickness of the viscous sublayer in the flow of
polymer solutions. But here it has to be considered that the commercial pipes have a non-
uniform distribution of roughness height and location of the roughness elements on the
pipe surface. The highest elements contribute the most to the modification of the turbulent
flow. It is therefore reasonable that roughness elements in the viscous sublayer produce
disturbances that can modify the turbulence in the polymer solution case. These
disturbances are broken up and disappear in Newtonian fluid flow, whereas in polymer
solutions they have a longer life time.
In practical applications of drag reduction this effect has to be taken into account. As a
consequence, in commercial pipes the non-uniform roughness influences the friction
behaviour of drag reducing polymer solutions at lower Reynolds numbers to a greater
extent than the Newtonian fluids.

5.5 Theories of drag reduction, or discussion of hypotheses

As already stated in chapter 4.3, the effect can be thought of as a result of a variety of
interactions of the additives with the flow. In this chapter we present some of the theories
based on ideas sketched in 4.3. However, as all these theories are not completely
satisfactory in view of all experimental evidences, this chapter contains a discussion of
theoretical hypotheses or possible mechanisms rather than a closed theory.
Let's start with a first statement. Whatever the mechanism may be, only elongated
molecules produce drag reduction. The dissolved polymer molecules are spongy spheres,
and can become elongated by frictional local force gradients produced by flow fields
acting on them. The elongated polymers can relax under the influence of Brownian
motion if the strength of the local flow field (IVlll~A. with A. given by eq. (2.4a)) has
decreased at a certain limiting value. The first question which therefore arises is whether
the molecules in their elongated form have to be present at any time in the whole flow
field when drag reduction occurs or whether it is sufficient that they become elongated
when the local flow field satisfies a criterion of strength?
134 CHAPTERV

In the first case one can argue that the flow should be described using a homogeneous
fluid with a non-Newtonian rheology. In the second case all kinds of micro-mechanical
interactions can be thought to be responsible for the effect. In the most simple model the
interaction of the polymers is related to the size of the structures in Newtonian turbulence.
The scales of turbulence are given by the size distribution of vortices present in the flow.
However, as we have seen in the topological representation, chapter 5.2.2.2, this is too
much of a simplification, since as we know, the structures with the highest efficiency for
stretching the molecules are the elongational parts of the flow field. These structures are
not all comprehended by a vortical analysis since their locations can be outside the
vortical structures.
In both cases the effect can be the result of the interaction of an ensemble of polymers
with the flow structures or of individual molecules. From dimensional consideration (a
fully stretched molecule has a length of about I0-5 m, see Fig. 1.1) the latter cannot be
the case. Drag reduction is, therefore, the result of neighbouring molecules interacting as
an ensemble of additives similarly embedded in an area of the local flow field.
Nevertheless, we will begin with a description based on alterations in the scales of the
vortical structures of the flow field.

5.5.1 The introduction of new length and time scales by the additives

Lumley (1969 & 1977) showed that the drag reducing effect by additives can be roughly
explained by dimensional reasoning.
The mechanism Lumley postulates is that under pure strain, the molecules are pulled into
an ellipsoidal shape. When the strain rate exceeds the restoring forces, the molecule
expands as far as it can until it is essentially fully expanded. If the molecule is in a shear
so that it is rotating as well as being subjected to the straining, a larger strain rate is
required to produce the same effect since the molecule remains aligned with the principal
axis of strain rate for a shorter time. Lumley's idea was that at high Reynolds numbers,
vorticity and strain rate are not correlated with one another and can occur in virtually any
combination. They both have distributions with long tails so that large and small values
of each are more probable than for a Gaussian distribution. Hence, there is a substantial
probability of finding regions of relatively large strain rate and relatively small vorticity,
giving rise to a change of the shear viscosity in these regions, see chap. 5.2.2.2. If shear
and strain rate are present simultaneously or with a small time shift, Lumley's assumption
may still be correct since, from experimental results, it is additionally known that the drag
reducing effect can be enhanced if shear and strain zones interact with a small time shift.
This is called the preshearing effect. The general idea, therefore, was to investigate in
which flow zones the polymer are deformed, thereby giving rise to a change in shear
DRAG REDUCTION IN POLYMER SOLUTIONS 135

viscosity. By this proceedure he found that the only difference in the boundary layer
turbulence structure between Newtonian flows and drag reducing flows is the scaling of
the various regions. Polymer molecules are expanded in the flow outside the viscous
sublayer due to the fluctuating strain rate. This causes an increase in the effective
viscosity, which in tum damps small dissipative eddies. Because the molecules are not
expanded in the viscous sublayer, the viscous sublayer is not affected by the polymers.
Together this results in a delay in the reduction of profile slope to a point farther from the
wall, and a consequent thickening of the wall region. Lumley argued that based on such
an analysis, it is the intensity of the smallest eddies which will be reduced and which can
explain the various aspects of drag reduction.
The large scale structures are uneffected by the polymers, therefore the main change in
the mean velocity profile is due to a change of the buffer zone resp. elastic zone.
If there is a range of y such that all the arguments off are either <<1 or >>1, the f will be
a constant

(_1_) au = const.
u't ()y (5.19)

and it can be assumed that the local mean velocities in the drag reducing flow would scale
with the friction velocity, as in the original flow, where the velocity profile in this range
of y is a logarithmic one

~=Ainyu't +B
u't v (5.20)

The constant may be determined by the reasoning based on the energy budget of the large
eddies, Tennekes (1966). Eq. (5.20) is well established experimentally, and it can be
concluded that the energy-containing eddies still scale with y. The constant, A, in eq.
(5.20) would be the same as that for Newtonian fluids if the solution had a linear relation
between stress and steady strain rate in simple shear. The fact that the velocity profile is
logarithmic is not so much Newtonian as inertial; the value of the slope is tied directly to
the constitutive relation in the sublayer. Hence, a logarithmic profile in some range of y
with the same slope as for Newtonian fluids can be expected, a result confirmed by the
experiments and schematically shown in Fig. 1.4.
We know that in the logarithmic layer there is no relevant length other than y. The peak of
the spectrum may be expected to occur at the wave number, k, given by
136 CHAP1ER V

(5.21)

which then represent the limit of the production range of the spectrum. In a Newtonian
boundary layer, the turbulence exists between a wave number, as given by eq. (5.21) and
a wave number defined by the peak in the dissipation spectrum which occurs for
Newtonian turbulence at

(5.22)

where 11 is the equilibrium range length-scale given by

-(v3)4
1

Tt- -
E
(5.23)

and E is here the production in the approximation, eq. (4.45). This is equivalent to the
Kolmogorov length scale, eq. (4.73). From the scaling of u'v' and 'iJu/()y it follows

( u3I
==> A- with
u== u't)
l=y
(5.24)

and eq. (5.23) with eq.(5.24) becomes

(5.25)

Hence, the peak of the dissipation spectrum would occur when introducing eq. (5.25)
into eq. (5.22) at

(5.26)

Lumley introduced the diagram Fig. 5.17 to show the interaction of the relevant length
scales in a turbulent flow. The area between the two lines represented by eqs. (5.21 &
5.26) describes the equilibrium range of turbulence in a pipe flow. The intercept of these
two boundaries, where production and dissipation become equal, occurs at y+ = 6.35.
DRAG REDUCTION IN POLYMER SOLUTIONS 137

This location roughly corresponds to the edge of the viscous sublayer. At this point
where viscous and inertial scales become identical, the turbulence becomes essentially
dissipative in character. Closer to the wall it cannot sustain itself.

log(y+) 102

Figure 5.17 Scaling relation in the constant-stress region of a turbulent boundary


layer, k+ being the non-dimensional wave number, Lumley (1969)

The ratio of a new length scale, A.+ 0 , to the length scale of a fixed eddy, given by its
wave number, k+, is

(5.27)

This relation appears in Fig. 5.17 as a vertical line, dashed line. As the wall shear
increases, A.n+ increases and k+ decreases. The vertical line, therefore, will shift to the
left. The region to the right of the viscous cutoff line is dominated by viscosity and
cannot be influenced by changes in turbulent scales. If k+A.n + = const. falls into the
equilibrium scale, then the boundaries of the dashed area in Fig. 5.17 are intersected by
this vertical in two points, A and B. Above y+=B the flow is unaffected; between y+=A
and y+=B the dissipative eddy scales are affected; and below y+=A the energy containing
eddies are affected.
138 CHAPTERV

A thickened sublayer and, consequently, drag reduction occur if the intercept of the
production and dissipation limits is shifted further away from the wall. Since the
production limit is assumed to be unaffected, the dissipation cutoff line has to be shifted
to the left. The sublayer thickening would be limited to the value of y+, corresponding to
A, since only the eddy scales to the right of the vertical line are affected. In other words,
Lumley kept the viscosity in the sublayer at its Newtonian value, but he assumed an
increased viscosity, VI, in the turbulent region. This implied that at any distance, y, from
the wall the viscous cutoff was shifted by the same amount in the log-log plot, Fig. 5.17.
The new viscous limit is represented by the bold full line parallel to the original limit,
with a slope according to eq. (5.22). The higher the concentration, the higher the shift of
the viscous limits. The net result is a shrinking of the turbulent domain. Beyond the
sublayer we now find a buffer layer increasing with the polymer concentration. It is
obvious then to expect that the turbulent losses will be reduced, Lumley (1973). The
whole effect occurs at arbitrarily low concentrations. For higher concentrations the
viscous limit shifts due to an increase in shear viscosity of the polymer solution. The
reason for this shift could be of diverse origins, like a violation of the Reynolds similarity
or a dependence on the new ratios of the length scales, etc.
This framework is too simple, however, to explain the detailed dynamics occuring in the
boundary layer. For example, it does not explain during which part of the turbulent
motion outside the sub layer the molecules expand, and how the expanded molecules are
distributed. Especially with this explanation the onset behaviour of the effect is not yet
incorporated. A theory would be needed relating the elongational state of the molecules
with the size of the affected dissipative eddies. As discussed by Lumley (1969), there are
many theoretical reasons to expect the time scale ratio, and none in favor of the length
scale ratio. It is, therefore, tentative to do this by relating the onset to the transient
behaviour of the polymers, or in other words, to their relaxation time. Lumley (1969),
therefore, transposed his length scale model into a time scale representation. The time
scale of a turbulent eddy is

'tk = ( k 2E )-1/3 (5.28)

so that a constant ratio of eddy time to 'tt. the time scale of the largest eddy, is given by

(5.29)
DRAG REDUCTION IN POLYMER SOLUTIONS 139

This is the line with a positive slope, shown in Fig 5.18. The reasoning is the same as for
the length scale ratio and the significance of the lettered points are the same.

104
I
I viscous
10 3 I dissipation
range

log(y+) 10 2

Figure 5.18 Scaling relation in the constant stress region of a turbulent boundary
layer. Turbulence to the right of the line with positive slope has a time scale
shorter than a fixed time

Again, the increase in viscosity in the turbulent part, and not in the viscous sublayer,
suppresses the eddies which carry the Reynolds stress in the buffer layer. This results in
the thickening of the sublayer and a reduction of drag. In agreement with this model the
large eddies remain essentially the same as in a Newtonian flow, but the scale of the large
eddies has expanded to fit the expanded sublayer. This is consistent with Lumley's
(1971) prediction that the large eddy structures are parasitic and primarily determined by
the mean velocity profile, and it is also consistent with several experimental results. The
large eddies lose their energy primarily at the bottom of the sublayer, where the molecules
are contracted; the viscosity is essentially unchanged from the Newtonian value. The big
eddies extend further away from the wall, the net velocity difference over the height is
larger, and the velocity fluctuations induced in stream wise direction are larger. Hence, the
peak value of the stream wise fluctuating velocity scales up as the sublayer increases and
is shifted. These are both observable results, see Fig. 5.4. ~he suppression of the small
eddies can be observed in the power spectra, Fig.5.13.
140 CHAPTERV

With this model Lumley (1977) explains the saturation of the effect with increasing
concentration as a feedback process. As the molecules are expanded by the strain rate it is
reduced because the dissipation for a constant v must remain constant. The reduction in
the strain rate stops the expansion. The equilibrium state is one of partially expanded
molecules, producing an increase in viscosity sufficient enough to hold the molecules at
that expansion. If more polymer is added, the average expansion is reduced, and the
effect remains about the same.
Lumley's model introduces a renormalized viscosity, which can be interpreted by
molecular mechanisms. However, it is known that flexible coils, even in the dilute
regime, behave elastically at high frequencies. De Gennes (1990), therefore, proposed a
model which makes use of elasticity arguments for the solution.
He used the properties of homogeneous, isotropic, three-dimensional turbulence without
any wall effect in the presence of polymer additives. The central idea of his model is to
show that at small scales (high frequencies) the cascade cannot be described by a
viscosity, but rather by an elastic modulus. The results of this investigation are then used
to describe the interaction of the polymer molecules with the wall turbulence, i.e. to set
up a modified version of Lumley's approach, where, at each distance from the wall, we
have an elastically truncated cascade. This gives a law for minimum eddy size, rmin.
versus distance, y, which is qualitatively different from Lumley's viscous effect. But the
net result is still an enhancement of the buffer layer.
De Gennes derived the idea for this kind of representation from results found by early
heterogeneous drag reducing experiments which seemed to be independent of wall
effects. As we showed in chapter 5.3, this is not the case. However, de Gennes's ideas
are still helpful for a better understanding of the effect.
To understand the difference between the two models, Lumley's and de Gennes's time
criteria have to be explained. At two points in the flow field separated by a distance of r,
the characteristic fluctuating velocity, u'(r), is related to the dissipation by the condition

u'3
-=E
r (5.30)

where u'/r defines the characteristic frequency (the hydrodynamic frequency) for this
system. This value has to be compared with the Zimm relaxation rate of a coil

(5.31)
DRAG REDUCTION IN POLYMER SOLUTIONS 141

with k being the Boltzmann constant , T the temperature and TJO the solvent viscosity.
Flory's law, eq. (2.10), holds for a linear, flexible, neutral polymer in good solvent for
RH.
At large scales, r, the hydrodynamic frequency is smaller than l{fz, but for a certain
value of r, l"='Tf, the two frequencies are equal and thus

u'(rr )Tz = rr (5.32)

With eqs. (5.30) and (5.32) rc can be evaluated as

rr = (ETz3)1/2 (5.33a)
and

'I' )1/2
uI r= ( Ez (5.33b)

where rc and u' f depend on RH and therefore, eq. (2.10), on the molecular weight and not
on the concentration.The condition, eq. (5.32), is the natural expression of Lumley's
time criterion, Lumley (1973). The interesting viscoelastic effects will occur only at
frequencies higher than l!fz, or equivalently at scales r<rc.
The Kolmogorov cascade is always truncated by viscous dissipation of a scale, rz,
defined by

_ (R er )3/4
rr -
-
rz (5.34)

with

(5.35)

the Reynolds number of this critical state.


When the coil is stretched by a factor, A., a certain elastic energy is stored in each of the
chains. With

1 << 1 <<N215 (1 = 100) (5.36)


142 CHAPTER V

the free energy, Er, of a coil due to the elastic anharmonic energy per unit volume is

(5.37)

where G' has the dimension of a storage modulus, which is a measure of recoverable
strain, see chapt. 3, and is linear in concentration. Equivalently, the stress, "te, due to the
restoring forces for c/N springs/ cm3, is

'te = Er (5.38)

Whenever 1:e is much smaller than the Reynolds stress, pu2, the reaction of the polymer
on the flow, is negligible. However, if we go to smaller scales, A and 'te increase. Thus,
at a certain scale, re, the two stresses become equal

(5.39)

with re being the elastic limit, which is observable only if re>rz. Since re is dependent on
the concentration, there exists a concentration threshold, Cm, below which the polymer
should have no visible effect. Since Cm is a strongly decreasing function of N, it may
well be that for long chains Cm is extremely small and hardly detectable.
An upper limit of the effect is given when the polymers become fully elongated. From eq.
(5.36) it follows that the maximum value in A, Amax. describes this status

'\ = N2/5
'\ = 1\.max-
A (5.40)

With the assumption that there exists a power law between the polymer elongation and
the spatial scale of the form

A.(r) = (~ r (5.41)

it follows that

(5.42)

with rm being the stretching limit.


DRAG REDUCTION IN POLYMER SOLUTIONS 143

In Fig. 5.19, which is a log-log plot of the stresses as a function of rr/r, when A. gets
close to Amax. the stresses tend to diverge. This fixes rm. It is seen from this plot how the
elastic concept by de Gennes differs from the fully stretched molecular state concept by
Lumley. In other words, the elastic correction is necessary for solutions beyond a
threshold concentration.

log 't
For a Kolmogorov cascade
truncated at re . The molecules
are only partly stretched.

_ _ For totally stretched molecules

- elastic limit
-stretching limit

log rr
r

Figure 5.19 Reynolds stresses and molecular stresses in a Kolmogorov cascade


with a dilute polymer solution. The upper representation shows the truncation
due to elastic effects for molecules with partly stretched chains; the lower
one, for fully stretched molecules. The latter representation is shifted in
vertical direction to be more comprehensible

The predictions found with these methods are nearly independent of the .detailed
mechanism(s) which we assume for the particle flow interaction, and therefore arguments
stated here can be used to explain all kinds of drag reduction by additives.

5.5.2 Changes in the burst process, an introduced stability mechanism

The burst process is naturally strongly related to the scales. As already mentioned,
changes in the scales related to the burst mechanism were observed in polymer solutions.
In chapter 5.5.1 the changes in scales were described by dimensional arguments due to
144 CHAPTER V

viscous effects based on the size of the smallest vortices. In other words, drag reduction
was explained via the influence of the additives on the dissipation mechanism of the
turbulent flow. Now we will introduce a completely different view in which drag
reduction is explained via the influence of the additives on the turbulence production.
The main idea is that the bursts are the result of an instability process responsible for the
turbulence production. Landahl (1977) proposed the following conceptual model, which
is completely described in Bark's (his Ph.D student) thesis (1974).

5.5.2.1 Landahl's instability theory

While the approaches in chapter 5.5.1 are statistical, Landahl's is mechanistic. He applied
the classical hydrodynamic stability concept to the turbulent boundary layer and
developed a so-called two-scale model. He considered the turbulence to consist of a
coupled motion at two disparate scales: a primary one of large scale, of the order of the
boundary layer thickness, and a much smaller secondary one, of the order of the viscous
sublayer thickness. The small-scale secondary motion ride on the primary wave, and
under certain conditions grow to very large amplitude, leading to a breakdown of the
primary waves. The mechanism of drag reduction due to the elongation of the added
molecules stabilize the small scale field and, thereby, inhibit turbulence production.
Through the excitation of the mean shear flow, caused by the strong mixing, a bursting
motion produces decaying wavelike disturbances, propagating downstream. As v' and p'
decay, there will remain au' perturbation in form of a convected eddy. Shearing this
eddy by the mean flow will further intensify the shear and produce a thin internal shear
layer, which will develop an inflection point and, thus, become locally unstable,
producing the primary scale disturbances. While this is going on, the shear layer is
intermittently bombarded with large-scale propagating disturbances originating upstream
in other bursting events. These disturbances will give the group velocity of the small-
scale secondary instability waves a large-scale variation in space and time. When the
secondary group velocity becomes locally equal to the phase velocity of a primary wave-
like disturbance, the focusing of the secondary disturbance on the primary wave will lead
to a rapid build-up of secondary wave energy. The first strong secondary disturbance
would likely appear more or less as a half-wave in the form of a "kink" of the shear layer.
The shear layer will kink upward or downward, leading to either an ejection or a sweep.
If the molecules inhibit the secondary instability, this would lower the turbulent stress
production, and according to the model we would achieve drag reduction. In all cases
where there is clear experimental evidence of drag reduction, a strong secondary
stabilization is observable, statistically showing up as a damping of v', whereas u' seems
DRAG REDUCTION IN POLYMER SOLUTIONS 145

not to be influenced in its intensity. To complete this model, one has to explain the
stabilization process.
Since viscoelastic effects normally are slightly destabilizing, e.g. Zakin et al. (1977),
viscoelasticity on the basis of this model cannot be the key fluid property producing drag
reduction. A more likely candidate is the anisotropic stress caused by the extension of the
molecular coils or aggregates in a turbulent flow of sufficiently high wall stresses. If
stretched strongly by the mean shear flow, such structures will have an overall effect on
unsteady perturbations, qualitatively similar to that of almost rigid elongated particles
suspended in the fluid.
In a pure shear flow the particles will tumble, but if they are elongated, they will tend to
spend most of the time oriented in or near the mean-flow direction. Thus, according to
the model, the rheological effect of polymer additives would be qualitatively the same as
that of fibres. For suspensions of rigid rods, if they are slender and aligned in the mean,
Batchelor (1970) showed that very high values of elongational viscosity without any
substantial increase in shear viscosity can be obtained.
This model can be criticized due to the fact that the bursting mechanism is more or less
explained by a non-viscous process. However, so far it is one of the best descriptions of
a self-reproducing process of structures of a turbulent boundary layer and is in
accordance with the main experimental results. This model has the advantage that it
explains drag reduction by a change in all scales. It is a model which is based on changes
in the turbulent production and it is a model which explains the "damping" of v' and the
rather unaffected u' 's. It propagates a stabilization effect of the bursting due to two
effects: (I) the alignment of the stretched polymer in flow direction and (2) the
elongational viscosity caused by these additives. The model would also explain the
similarity of this drag reducing effect with those produced by surfactants and fibres.
Onset, in this case, occurs when the two conditions mentioned above are satisfied. The
only problem is that one has to assume that the turbulent core-flow is independent of the
kind of instability process close to the wall. On the other hand, the maximum drag
reduction asymptote can be explained with a state in which the two conditions hold for
the entire cross section.
Landahl ( 1990) extended his model of a burst cycle by a new mechanism which includes
an algebraic instability mechanism, leading to the subsequent creation and an evolution of
a streaky structure. But as far as drag reduction is concerned, the explanation remains the
same.
146 CHAPTERV

5.5.2.2 Stabilization of the bursting

The stabilization of the bursting mechanism based on a phenomenological model was


introduced by Smith et al. (1990). It is an intuitive model based on visualization of the
structures of the turbulent boundary layer, which are described mainly by hairpin vortices
and the product of their interaction. Smith et al. argued that the reduction of turbulent
surface drag requires a reduction in the level of momentum exchange at the wall, which in
tum requires a reduction of the bursting activity near the wall. In principle, this may be
accomplished either by reducing the number of low speed streaks adjacent to the wall or
by increasing the cycle time for the momentum exchange bursting process. The hairpin
vortices, this means a structure composed of two counterrotating longitudinal vortices
closed by a loop at the outer ends, produce the streaks and eventually provoke an
eruption to produce new hairpin vortices. A simple model of this kind can already be
found in Offen & Kline's paper (1975), or more recently in Perry & Chong's (1982)
work. It defines a cyclical momentum exchange process with the surface flow, which can
be interrupted in at least two ways. The first is to provide mechanisms which inhibit the
viscous-invicid interaction by interfering with the lateral movement of low-speed fluid at
the wall, as it is thought that riblets reduce drag, Bechert & Bartenwerfer (1990) or
Luchini et al. (1991). The second is to maintain the streaks in a stable state for a longer
period. This is equivalent to Landahl's assumption. The maintenance of streak stability is
a tenuous approach since it entails a delicate balance between the inherent stability of the
low-speed streak and the amplitude of the destabilizing pressure perturbation in the outer
flow. The addition of polymer to the boundary layer is an example whereby streak
stability is increased by the addition of an external additive, as evidenced by wider streak
spacing and reduced bursting activity. Smith et al. (1990) cannot give a theory for this
stabilization. They speculate, however, that the polymers either directly affect the streak
stability by inhibiting lateral concentration of fluid by the hairpin vortices, or indirectly by
providing a region which is locally more viscous. This would result in a more effective
damping of the external perturbations, thus preventing streak break-down. It would also
dissipate the energy in the hairpin vortices generated by breakdowns, thus weakening the
vortex strength of the hairpin vortices and inhibiting their effectiveness in the viscous-
invicid regeneration process. Based on their Overall Production Modul, as the main
feature of the turbulent boundary layer, Falco et al. (1990) suggested that drag can be
reduced by modifying the wall so as to reduce the amplification of strong vortical motions
created near the wall. Such descriptive "theories" may be a useful guideline for planning.
However, they are far from being a real theory, as is the case in Lumley's, de Gennes' or
Landahl's model.
DRAG REDUCTION IN POLYMER SOLUTIONS 147

5.5.3 Development of anisotropic low dissipative structures

The structural models used to explain drag reduction, based on so-called "coherent"
structures, were rather crude. As mentioned in chapter 5.2.2.1, the structures in a
turbulent boundary layer flow are much richer and therefore also statistically, rather than
mechanistically describable.
Based on this new concept of a topological representation of the flow field, the question
arises, how do the polymers interact with the various typical patterns. The idea of such a
local interaction with "damping" effects stems from the fact that drag reduction is
associated with various mechanisms of rheological nature which operate at small scales.
Dilute polymer solutions can, however, significantly influence the large scale structures.
An example is Riediger's (1988) observation of a mixing layer in which 320 mm
downstream of the splitter plate, the structures in water were nearly destroyed, whereas
in the polymer solution the flow field showed the characteristic shape for large vortices of
the plain mixing layer. This is caused by a stronger dissipation of energy in the water.
Thus, the explanation must contain a mechanism influencing all scales of the flow field.
Tsinober (1990) stresses the fact that drag reduction is generally not associated with
reduction of turbulent intensity. It may not be necessary to suppress the turbulence or
even its production, but it must be sufficient to let the flow remain turbulent, "somehow"
altering its structure so as to reduce turbulent dissipation of energy and, thereby, the
turbulent drag.
We have already mentioned an attempt along such lines. Based on the idea that the
elongational viscosity of the particles is the reason for the effect, it has been checked,
with rather negative results, den Toonder et a1.(1994), by investigating if a suppression
or damping of the local strain fields could be responsible for the effect. This is a key
result as it shows that the act of stretching the polymers on the flow is not as important as
the stretching of the molecules itself. Elongated molecules are much easier aligned, and
this seems the most important contribution to drag reduction. In other words, first the
right particles have to be produced and then the effect can take place in a rather universal
form. Thus, it is assumed that the main part of drag reduction is due to a universal
mechanism based on the alignment of the elongated additives in local flow direction,
naturally with some specific differences due to the properties of the particles. Due to their
flexibility, polymer molecules will be rather ineffective in flows with dominant shear-like
structures, but very efficient within structures with curved flow lines.
How can such a universal mechanism be understood in the framework of geometrical
terms, as are introduced by a topological description of the flow field? In other words,
the question is not whether to stop the turbulence production, but rather whether we can
148 CHAPTERV

influence turbulence so as to be only slightly dissipative. Tsinober (1990) suggested


attributing drag reduction to the development of anisotropic flow structure. In other
words, the additives change the structure so that they belong to a class which is less
dissipative than the Newtonian case. The idea that anisotropy may be the key to this effect
has two roots. The first one is that the main change in drag reduction is the change from
isotropy to anisotropy in the JPDF of u'v', as shown in Fig. 5.12, and the second is that
magnetic fields, whose change Tsinober showed is due to a change in isotropy of the
velocity fluctuations, Tsinober (1990 & 1990a), are drag reducing too.
The argumentation is that those flow structures are low-dissipative which suppress the
process of the enstrophy, eq. (4.51), generation , eq. (4.97) - the only mechanism
sustaining a high level of turbulent dissipation.
Two idealized cases of minimum enstrophy production are known:
(1) two-dimensional turbulent flows, i.e., flows in which the instantaneous structure
depends on two Cartesian coordinates only.
(2) the Beltrami flows, i.e., flows in which mis parallel to y.

[!"!!.!!] = 0 (5.43)

In both, the generation process of mean enstrophy, see eq. (4.57), does not exist, i.e.,

(5.44)

and here the energy dissipation, eq. (4.54), can be expected to be very small.
The answer to the question, which class of anisotropic flows is low or non-dissipative,
is: the flow has to have a low value in the enstrophy production or, in the extreme, it has
to fulfill eq. (5.44). Since

(roirokdik) =([!!!,rot[!!!,!!]])= (([m,!!],rotm)) = (([m,rotm],!!)) =


=(([!!,rot!!!]),!!!) (5.45)

the general condition is that one of<...> = 0 or the expression inside the brackets should
be a divergence, i.e.,

(5.46)
........ ,.etc.
The general structure of fields satisfying these conditions is unknown. However, a broad
class can be found as a consequence of solenoidality of vector .1! and m, e.g.
DRAG REDUCTION IN POLYMER SOLUTIONS 149

[!!!,!!] = grada (5.47a)


[!!!,rOt!!!] = gradf3 (5.47b)
[!!,rOt!!!]= grady (5.47c)
where a,J3;y are some scalar fields.
The two-dimensional flows satisfy the condition (5.47b) since

(5.48)

As Tsinober suggested, these three relations can be interpreted as constraints imposed on


the interaction between energy containing eddies and eddies of the order of the Taylor
microscale, eq. (4.70), and between eddies of Taylor microscale and eddies of the order
of the Kolmogoroff scale, eq. (4.73 & 4.74), respectively. Each of these constraints can
be effective in drag reduction.
The left side of eq. (5.47a) is the definition of the Lamb-vector

(5.49)

and, therefore, the Beltrami flows, eq. (5.43), are the special case for which the Lamb-
vector fulfills condition (5.47a) and is a potential vector.
With this in mind, one can raise the question whether in drag reducing flows an increase
of the potential part of the Lamb-vector can be observed. This implies that we can
measme A. and its change with respect to the Newtonian case.
Moffatt (1989) suggested that turbulent flows in general may spend large portions of their
time in a neighbourhood of fixed points of the Euler equation. This implies that in
turbulent flows the potential part of the Lamb-vector can be expected to be larger than its
solenoidal part. Tsinober (1989) analytically showed that the Lamb-vector for a random
Gaussian solenoidal vector field, .1!. is 70% potential, in the sense that 70% of the mean
square, <A.2>, are contributed by its potential part. The first implication of this result is
that there is a strong reduction of nonlinearity in turbulent flows due to purely kinematic
constraints. Another aspect is that, again for purely kinematic reasons, the turbulent flow
is imposed on certain constraints in its topology, i.e., vectors !ll andy are close tq tangent
to a-sufaces, where a is the potential of the potential part of the Lamb-vector. It is the
solenoidal part of ~ which should be suppressed in order to achieve drag reduction
effects since only this part contributes to the enstrophy generation because
150 CHAPTERV

( ro'i ro' j dij) = -( (rot~, rot~))


(5.50)

where rot ~ is the solenoidal part of f..


This is equivalent to a flow field with a spotty distribution of the helicity density, eq.
(4.25), that means areas of high helicity and, therefore, of low dissipation. The alignment
of the molecules produces local and instantaneous anisotropies of the rheology of the
local flow field, in the sense that they enhance the helical densities in the "hot spots" of
turbulence. If eq. (5.43) holds, then the helicity density becomes a maximum. In other
words, the change in helicity density and in the potential part of A. are related.
Although these are highly hypothetical remarks, it seems that these topological constraints
are very promising in unifying the theoretical descriptions for all elongated drag reducing
additives and probably also of the passive drag reducing phenomena.

5.5.4 Virk's type B drag reduction

De Gennes discerns between the randomly coiled molecules, which are deformed by the
stretching due to the flow into ellipsoidal spongy balls, and the fully extended molecules,
as used by Lumley for his theoretical explanation. Virk & Wagger (1990) used the same
properties together with the main features of the "coherent structure" to classify drag
reduction mechanisms by an additive-burst matrix (ABM) for drag reducing
macromolecules, Fig. 5.21.
By type A drag reduction Virk & Wagger mean a family of additive solutions yielding to
friction factor segments fanning outwards from a common "onset" point on the Prandtl-
Karman line, see Fig. 5.20, their slope increasing with increasing concentration, and
drag reduction increasing with increasing Re--ff. Virk & Wagger postulate that this is
characteristic of random-coiling macromolecules undergoing a transition due to
stretching.
In type B drag reduction a family of additive solutions yields on Fig. 5.20 segments that
are roughly parallel to, but displaced upwards from, the P-K line, with drag reduction
essentially independent of Re-Jf, but increasing with increasing concentration. This
behaviour is exhibited by a variety of additives, including extended polyelectrolyes,
fibres, soaps, and clays.
Fig. 5.20 schematically shows the main features of these two types of drag reduction.
Fairly truncated or slightly extended polymers, produced by added salt, behave as type A
drag reducers, while very extended polymers behave as type B drag reducers. Using
partially deformed molecules, the laminar drag increases only a little bit; using extended
molecules it increases and quite appreciably due to the larger hydrodynamic volume. The
DRAG REDUCTION IN POLYMER SOLUTIONS 151

progressively larger drag reduction efficacies of the more extended conformation imply
that they must ever more closely resemble the fibre like state, which ultimately interact
with turbulence.

25r-----------------------~----------~

fl/2
Type B.

...
....
3
20

I
.
1 Type A
~~ I
I
1 1 h

......
~~ I
15 I
I I
I

..
I
I
~~I
... ... t~.,.;
.-.. ' Ill
I/ .: .... ......
10 2
.:
.:3

5
2 2.5 3.0 3.5
1/2
log10 (Re f )

Figure 5.20 Scheme of the friction law for additives producing type A and B drag
reduction. (L) laminar behaviour, (N) Newtonian behaviour for smooth
walls, P-K law, (M) at maximum drag reduction. 0 stands for the onset
point, Or stands for its counterpart, the retro-onset point

For type A the friction law approximately follows the laminar curve and then falls onto
the P-K line, from which the curve restarts at an onset point with a fan pattern and with a
o,
slope increment, depending presumably on the extension of the molecules.
For the type B the onset character is lost. This is of utmost importance since it shows that
drag reduction does not need an onset; on the contrary, such solutions show a retro-onset
in so far as the onset is now on the M-curve and falls back into the sector between M and
N, with a slope normally parallel to the P-K curve. The onset phenomenon is, therefore,
152 CHAPTERV

related to a minimal extension of the molecules needed, whereas when the molecules are
extended, the mechanism is more general. The latter case has to do with the ability of the
additives to align in local flow direction if the additive-pervaded volume fraction is X v
>> 1 with an interparticle separation smaller than their extended length, Le. during drag
reduction. Xv is given by Virk & Wagger (1990), as

(5.51a)

for macromolecules or for fibres as

(5.51b)

Where Lc is the macromolecule contour length, N is Avogadro's number 6.021023, p


the solvent density, lr the fibril length, dr the fibril diameter, pr the fibre density.
Appropriate conversion factors were applied for length in mm, densities in g/cm3, and
concentrations in wppm.

Burst events lift-up growth breakdown

Type
Additive
states
-
--
-/\.
----
- 0
~0
u
u
__...u

~v,w

randomly coiled

* )
Is unlikely to be very important. Drag reduction
by spheres is invariably rather small
A

( )( )( )
coil
? ?
X
stretch +

transition
extended

J
I'

~
B Is almost certainly relevant to any drag reduction

'
Figure 5.21 Virk's Additive-Burst Matrix (ABM) for drag reduction by macromolecules
DRAG REDUCTION IN POLYMER SOLUTIONS !53

Assuming that drag reduction is mainly the result of an influence of the additives on the
bursting process, it is naturel to classify type A and B in terms of this influence. The
assumption itself is sustained by three kinds of evidence:
(1) At the onset of drag reduction, ratios of characteristic macromolecular-to-wall
turbulence length and time scales are invariant, and are closely related.
(2) The onset of drag reduction occurs at the same wall shear stress in hydro-
dynamically smooth and fully rough pipes, implying that the additive-turbulence
interaction commences outside the viscous sublayer.
(3) The mean velocity (and turbulent intensity) profiles during drag reduction show
that a characteristic, new region, the "elastic sublayer," arises in the vicinity of
y+= 15, where the turbulence energy production associated with bursts is at a
maximum.
In the ABM the object of the mechanistic search is to identify the matrix element(s) most
responsible for drag reduction.
The lift-up mechanism is associated with strong strain fields. Therefore it has to be
damped (X), see Fig. 5.21, whenever the molecules in this region can be elongated by
these strain fields. Growth and breakdown processes can be influenced only if the
molecules are already deformed in the lift-up process. In the ABM these processes are
marked with a question mark (?) to refer to two questions of particular interest already
encountered:
(1) Does the act of macromolecular elongation contribute to drag reduction? Or does
the latter depend solely on the fmal amount of elongated macromolecules?
(2) Which skeletal attributes govern the effectiveness of an extended additive?
From the ABM scheme one can conclude that type A drag reduction is the specific
polymer-like drag reduction, whereas the type B drag reducing mechanism(s) is common
to all efficient drag reducers. Therefore, type B drag reduction is the most interesting one.

S.S.S Some concluding remarks

The first of three different theoretical approaches was described in chap. 5.5.1-3, and
each of them begins with a different view of the interaction of the particles with the
turbulent flow structures. It would, therefore, seem that these three theoretical approaches
contradict one another.
In the Lumley or de Gennes model the polymers produce a new, but simple rheology,
which is mainly characterized by a change in viscosity. It is one, however, which is
effective only outside the viscous sublayer. The reason is that the change in viscosity is
due to the additives which need to be deformed by the flow, and this can be achieved
only by small scale structures or high frequency disturbances due to the short relaxation
154 CHAPTERV

time of the deformed molecules. Vice versa the deformed molecules suppress the small
structures.The interaction, therefore, changes the dissipation.
Landahl's model is based on stability arguments. Turbulence is thought to be dominated
by the so-called coherent structures. Since the scales of the structures produced by the
instabilities are influenced by the polymers, these flows are altered in their turbulence
production. It is a kind of turbulence suppression mechanism.
The third model is based on a change in all scales due to changes in the geometry of the
flow field. This is attributed to polymers which change the local and instantaneous flow
field by their alignment. According to this view, the polymers interact over the whole
cascade process, containing the production, the transfer and the dissipation of
turbulent energy.
It shall be stressed that the contradiction between the three models is an apparent one
rather than a real one. All three models show that if the polymer can change the rheology
locally, then the fluid can interact with the structure of the turbulent flow. In the small
scales of the flow this is always the case, and, therefore, the influence on the dissipation
is permanently present. In Landahl's approach the instability of the relevant wall-near
structure is damped. The reason for this damping is the stabilization of the secondary
wave instability by the fluid containing deformed polymers. Since these instabilities are
also of small scale, the basic idea in 'both models is very similar. The difference are: in the
first case the interaction is a direct one, whereas in the second the influence is due to a
feedback mechanism in which the production plays a dominant role. In the third model
the idea is that the topology of the turbulent flow field is so complex that the interaction is
present over the whole cascade process, since there are always locations in the flow field
where the additives can interact by local change in the rheology.
This comparison shows a very important aspect. Lumley or de Gennes introduce a new
rheology which is fairly homogeneous, whereas the two other models assume a very
spotty change in the rheology. In chapter 4.3 the implication of these differences was
discussed.
In view of the topological aspects discussed in chapter 5.2.2.2 it seems more realistic to
continue with a concept of local ("spotty") interaction between the flow field and the
particles. It would, therefore, be tempting to investigate the differences in the topology of
the flow fields with and without additives. This will soon become possible since such
developments in techniques like particle tracking have been made that the flow field can
be observed at high resolution at moderate Reynolds numbers. However, until now this
method has not been applied to investigations of the problem of drag reduction. Such
investigations would also help to verify Virk & Wagger's idea, namely, that the main
drag reducing mechanism is a universal one due to the local alignment of already
elongated molecules or analogous additives with a very large aspect ratio. According to
DRAG REDUCTION IN POLYMER SOLUTIONS !55

this view, the onset behaviour is necessary so that the system is fed with molecules large
enough, but the onset has practically no importance for the actual drag reduction
mechanism.
VI DRAG REDUCTION IN SURFACTANT SOLUTIONS

Although one of the first publications on drag reduction by additives deals with surfactant
solutions, Mysels (1949), these types of additives have received less attention than
others, especially polymers. However, in the past years this has changed because drag
reducing surfactant solutions do not degrade over long periods of time. This is a definite
advantage for practical applications, especially in closed flow circuits, and for studying
the structure of turbulence under drag reducing conditions because the rheological
solution properties and, consequently, the drag reduction do not change with time.
Furthermore, systematic research studies in the chemical industry, documented by several
European and U.S. patents, have led to the discovery of surfactant systems exhibiting
drag reduction at surfactant concentrations below 100 ppm.

6.1 Friction Behaviour

The turbulent friction behaviour of drag reducing surfactant solutions depends on the
size, number and surface charge of the micelles, see chapter 2.3 Since these properties
change with temperature and with the concentration of the surfactants and any additional
electrolytes, the friction behaviour of surfactant solutions is influenced in a similar way
by these properties. The main characteristic feature of the friction behaviour of surfactant
solutions is the disappearance of drag reduction when a critical wall shear stress is
exceeded. This critical wall shear stress can be up to 100 Pa, which is quite high in
comparison to the wall shear stresses where mechanical degradation starts in drag
reducing dilute polymer solutions. However, in contrast to polymer solutions, the loss in
drag reduction beyond the critical wall shear stress is reversible, i.e., drag reduction
recovers, when the shear stress is lowered again. The critical wall shear stress is
independent of the pipe diameter and depends on the properties mentioned above. Thus
drag reduction in surfactant solutions was also called a "wall shear stress controlled
phenomenon" by Savins (1967), one of the pioneers in surfactant drag reduction. By
using the method of small-angle-neutron-scattering (SANS), Bewersdorff et al. (1986)
demonstrated that the loss of drag reduction above the critical wall shear stress is
accompanied by a vanishing of the orientation of the micelles in the turbulent flow. The
micelles were not destroyed above the critical wall shear stress as stated earlier, but they
were disaligned. This is in agreement with micelle kinetics, because the time constants for
the formation of the rod-like micelles in the surfactant system used by Bewersdorff et al.
(1986) were in the order of 1Q3- 104 s. Thus, if the micelles would be destroyed above
the critical wall shear stress, a long time delay in drag reduction would be expected to
!57
!58 CHAPTER VI

occur when the critical wall shear stress is lowered below the critical value. This has not
been observed in experiments.
In a friction-factor-Reynolds-number plot the Reynolds number is normally, as is the case
for polymer solutions, based on the solvent viscosity. Without any doubt this is a
convenient plot, especially for practical applications, if one wants to know the drag
reduction at a constant flow-rate due to the dissolved additives. On the other hand, if one
wants to check whether the friction behaviour of drag reducing surfactant solutions is
governed by the same empirical friction laws as dilute polymer solutions, especially if the
drag is limited by the same asymptotic law, one faces severe problems when the shear
viscosity of the surfactant solution is noticeably increased in comparison to the solvent.
Furthermore, as shown in Chapter 3.5, the shear viscosity of these surfactant systems in
the shear induced state (SIS) can increase with the gap width of a Couette viscometer.
Thus, it becomes extremely difficult, if not impossible, to calculate a Reynolds number or
a dimensionless wall distance, y+, based on the solution viscosity for these surfactant
solutions.

f "' 6 .. .
A d=6 mm
d:50mm

..
I
2
6

.
A

~
A

.. .
--------6A"'

s--

"" " ~ :---------------- -----


A
6
6
A
AA

......
~ .__________

.
6

~
.".. . ...,.
6 "'
A A

~
~~
...
10~ 2

Re

Figure 6.1 Friction behaviour of an equimolar mixture of hexadecyltrimethylammonium-


bromide and sodiumsalicylate in tap water at T=25 C in pipes of different
diameters at c= 830 ppm

The friction behaviour of an equimolar mixture of hexadecyltrimethylammoniumbromide


and sodiumsalicylate in tap water, shown in Fig. 6.1, elucidates the influence of a shear
viscosity, considerably increased in comparison to the solvent, on the friction behaviour.
In Fig. 6.1 the Reynolds number is based on the viscosity of the solvent, i.e., water. At
DRAG REDUCTION IN SURFACTANT SOLUTIONS !59

low Reynolds numbers the friction factor decreases more steeply than predicted by
Poiseuille's law for a Newtonian fluid. In a laminar flow this could be due to shear
thinning occurring at low shear rates in this fluid. When the Reynolds number is
increased, the friction data intersect with the Prandtl-Karnuin law, eq. (1.12), i.e., at low
Reynolds numbers, these solutions exhibit a drag increase. Furthermore, at low Reynolds
numbers the deviation of the friction data from Poiseuille's law is larger in a 50 mm pipe
than in a 6 mm pipe. Thus the question remains whether the shear viscosity still increases
with the pipe diameter. The critical wall shear stress at the break-down of drag reduction
is 5.8 Pain the 6 mm pipe and 5.3 Pain the 50 mm pipe, which means it remains nearly
constant. In these experiments the break-down occurred before the friction data
approached Virk's maximum drag reduction asymptote, eq. (1.13). However, this may
also be due to the fact that these solutions exhibit remarkably increased shear viscosities
in comparison to the solvent in the complete shear rate range, and the Reynolds number is
based on the solvent viscosity.

-----
2

.~ ----.,____
~
(

7
~~

"'-~
~0 ... ~0
0
0
0
0
o o Oo
- -------..:
'
0
~
~ o o,.o~'b""

2 -
c-700 ppm
~
T=16"C
~
o
A T=18"C
o T=25"c
2 4 7 2 4

Re

Figure 6.2 Friction behaviour of an equimolar mixture of tetradecyltrimethylammonium-


bromide and sodium salicylate in de-ionized water for different temperaturs in
a 15 mm pipe at c= 700 ppm

In other surfactant systems, e.g. Chara et al. (1993), Bewersdorff & Thiel (1993), or
Beiersdorfer et al. (1994) the friction data intersect Virk's maximum drag reduction
asymptote, and in a certain Reynolds number region, exhibited drag reduction which was
larger than predicted by eq. (1.13). This even occurred in artificially roughened pipes,
160 CHAPTER VI

Bewersdorff & Thiel (1993). They found no influence of the surface finish on drag
reduction at low Reynolds numbers (Re< 30 000), which is in agreement with the results
for dilute polymer solutions reported by Virk (1971), who found an identical friction
behaviour in smooth and rough pipes when the dimensionless roughness height was
k+<l2, see chap. 5.4. In the Reynolds number region, where the drag reduction exceeds
Virk's maximum drag reduction asymptote, the flow was turbulent.
Due to the influence of the increased shear viscosity on the shape of the curves in a
friction-factor-Reynolds-number plot, solutions of equimolar mixtures of tetradecyl-
trimethylammoniumbromide and sodium salicylate show different shapes of their friction
data. The shear viscosity of these dilute solutions below the occurrence of the shear
induced state (SIS) is only slightly increased in comparison to the Newtonian solvent.
Fig. 6.2 shows the friction behaviour of such a solution at three temperatures in a 15 mm
pipe. Due to the limit in the maximum flow rate of the pump in the flow circuit, the break-
down of drag reduction could not be reached. At low Reynolds numbers the friction data
in Fig. 6.2 follow Poiseuille's law of a laminar pipe flow. For the two lower
temperatures the data follow the laminar curve up toRe= 4,000- 5,000 before deviating
to Virk's maximum drag reduction asymptote. Therefore, the transition laminar-turbulent
is shifted to higher Reynolds numbers. This observed shift of the transition is in
agreement with stability calculations for stiff rods, as done by Gyr (1976) and Bark &
Tinoco (1978). At T = 25 Oc the surfactant solution exhibits an onset phenomenon, i.e.,
deviations from the Newtonian turbulent friction behaviour only occur when a critical
Reynolds number is exceeded. Experiments with an observed onset behaviour, well
known for dilute polymer solutions, can be used to check the influence of the SIS on drag
reduction if, in addition, shear viscosity measurements are done.
Fig. 6.3 shows the friction behaviour of equimolar mixtures of tetradecyltrimethyl-
ammoniumbromide and sodiumsalicylate in a 15 mm pipe. To produce this onset
behaviour in the turbulent flow regime, 2 mmol/1 sodiumbromide were added as an
additional electrolyte. By assuming that at the onset-point the turbulent velocity profiles of
a surfactant solution and a Newtonian fluid do not differ, the turbulent velocity profile of
a Newtonian fluid can be used to calculate the dimensionless wall distance, y+, where the
critical shear gradient for the SIS occurs. At onset conditions this critical shear gradient is
found in the buffer zone (y+ = 4.7- 6.4). This means that drag reduction in surfactant
solutions occurs when shear induced structures can be built up in the buffer zone. In
many cases the friction behaviour of surfactant solutions shows no onset behaviour, and
the friction factor follows Poiseuille's law in the laminar flow regime and an equivalent to
Virk's maximum drag reduction asymptote in the turbulent flow regime. In these cases
the critical shear rate for the occurrence of the SIS is already exceeded in the laminar flow
regime.
DRAG REDUCTION IN SURFACTANT SOLUTIONS 161

2
f

.~ ---.,...,._

"
10
D 0~
'-q_....._"" 0 "'
~ ....,._
7 D ...._
"'
~ ~
0

... D D D

""
0
Deco
0 ......
...... 0

.......... 0 0
c-1000 ppm o oo ooooo
c....,.-2 mmol/1 ~
2 f- 0 T=2oc
~ .........

0
"
T=2sc
T=3oc
~
7 to 4

Re

Figure 6.3 Friction behaviour of an equimolar mixture of tetradecyltrimethylammonium-


bromide and sodiumsalicylate in de-ionized water for different temperatures in
a 15 mm pipe at c=lOOO ppm with an additional electrolyte concentration of
CNaBr= 2mmoVI

SANS-experiments by Bewersdorff et al. (1989) and Lindner et al.(1990) showed that no


anisotropies in the scattering pattern existed below the onset of drag reduction, i.e., the
micelles were statistically oriented in the flow. After the onset of drag reduction the
anisotropy of the scattering pattern was correlated with the drag reduction, and when the
friction factor approached a maximum drag reduction asymptote, a nearly complete
alignment of the rod-like micelles in flow direction from the scattering pattern was
detected. The SANS-experiments demonstrate that the micelles are oriented with their
long axis parallel to the flow direction in laminar Couette as well as in turbulent drag
reduced pipe flows when a critical shear gradient is exceeded.

6.2 Velocity profile and structure of turbulence

Results of turbulent velocity profile measurements of drag reducing surfactant solutions


are reported in both channel (Povkh et al. (1975, 1980, 1988), Beiersdorfer et al. (1994))
and pipe flows (Bewersdorff & Ohlendorf (1988), Bewersdorff (1990), Chara et al.
(1993)). In general, the viscous sublayer seems to remain unchanged in comparison to a
Newtonian fluid, whereas in the other parts of the velocity profile the shape of the profile
162 CHAPTER VI

exhibits no general tendency. This may be due to several reasons. Usually the solvent
viscosity is used to calculate the dimensionless wall distance, y+. However, depending
on the surfactant system, the surfactant and additional electrolyte concentration, and the
temperature surfactant solutions can exhibit large increases in the shear viscosity in
comparison to the solvent, see Chapter 3, which additionally depend on the shear rate.
Thus, in a laminar pipe flow the shear rate would vary with the wall distance. In turbulent
pipe flows of drag reducing surfactant solutions the actual effective viscosity which the
solution exhibits is velocity field dependent. To make things more complex, one has
additionally to consider the influence of the time effects, i.e., the influence of the
deformation history, on the actual viscosity.

Figure 6.4 Dimensionless velocity profiles of dilute surfactant solutions for different
concentrations of "Methaupon" at different pressure drops in a rectangular
channel (20X50 mm), from Povkh et al. (1980)

Fig. 6.4 presents results of velocity profile measurements in a rectangular channel for
different concentrations of a surfactant called "Methaupon," from Povkh et al. (1980). In
these velocity profiles the solvent viscosity was used to calculate y+. The velocity data
follow at the end of the viscous sublayer, which seems to reach up toy+ = 11, as found
in the flow of drag reducing polymer solutions, an ultimate profile in the buffer zone. In
the core region some velocity profiles exhibit a parallel shift in comparison to a
Newtonian fluid, as found for drag reducing dilute polymer solutions, whereas the other
DRAG REDUCTION IN SURFACTANT SOLUTIONS 163

profiles show an increased slope in this region, which can be due to a change in the
structure of turbulence or due to a viscosity, varying with the wall distance (which
would, of course, also influence the structure of turbulence).
When the friction factor lies close to Virk's maximum drag reduction asymptote, the
velocity profiles in the buffer layer become S-shaped, i.e., they deviate from Virk's
ultimate profile to lower dimensionless velocity values at smaller wall distances, then to
higher ones in the medium range, and again towards lower ones at larger wall distances.
TheseS-shaped velocity profiles were also found by Bewersdorff & Ohlendorf (1988),
Bewersdorff (1990), and by Chara et al. (1993). The shape of these dimensionless
velocity profiles at maximum drag reduction conditions rather resembles a laminar than a
logarithmic turbulent profile.

70
Re= 9600
60 x Re=l2500
* Re=ISOOO
o water
40

20
... ~~
._______,________ : ________ I
0
1 10 100 1000

Figure 6.5 Dimensionless velocity profiles of a solution of an equimolar mixture of


tetradecyltrimethylammoniumbromide and sodiumsalicylate at c= 1.9 mmol/1
in a quadratic channel, a= 15 mm

Beiersdorfer (1994), Beiersdorfer et al. (1994) used the viscosity of the surfactant
solution at low shear rates, below the shear induced state, for calculating the
dimensionless wall distance, y+. Their results, see Fig. 6.5, show velocity profiles
differing in shape from those of Newtonian fluids as well-as from those of drag reducing
polymer solutions. At the end of the viscous sublayer, which reaches up to y+ = 5, as
164 CHAPTER VI

found for Newtonian fluids, a logarithmic velocity profile is found in the buffer layer
which is identical for different Reynols numbers

u+ = 15.251ny+ -20.25 (6.1)

Chara et al. (1993), who used the solvent viscosity to calculate y+, found for their
velocity profiles in the buffer layer by fitting their data

u+ = 23.4llny+ -65 (6.2)

50,---~-r,--,-.,----r-,-,--,-,---,--,~~~

I ~ I R~.S~~ u.,1\71n~-1Vr
I : v/ I
I/
'1,

" .. rTi tr- ~I - t;vv~1 ~r-'rT


'I

20
II II
---f--
II
~- ---
j-J:4 "
.,P
~
0
0
o
o-~l-1----


l

I
.;

!
water
I I
I I

I I _./ " c/'/c ~ ,. "'


D
s~lvcn_t viscosity
VISCOSity at 1:w
1

iI
.. ~ "' Variable viscosity

a ............ ~-- -9 .. u>, ~ I II II


~ 6 8~ 6 8~

Figure 6.6 Influence of the behaviour of the shear viscosity on the shape of a
dimensionless velocity profile of an equimolar mixture of hexadecyltrimethyl-
ammoniumbromide and sodium salicylate at c = 830 ppm. The dimensionless
wall distance, y+, was calculated by using the solvent viscosity, the solution
viscosity at the wall shear stress, and a variable viscosity according to eq.
(6.3)

Thus, it becomes evident that Virk's ultimate profile, eq. (1.16), for drag reducing
polymer solutions is no more valid for drag reducing surfactant solutions. The limit
between the buffer layer and the core region is shifted to higher y+-values with increasing
DRAG REDUCTION IN SURFACTANT SOLUTIONS 165

Reynolds number in Fig. 6.5. However, this deviation from eq. (6.1) always occurs at
the same absolute wall distance ofy = 4 mm. Consequently, this change in the structure
of turbulence may occur at a fixed wall distance and does not depend on the Reynolds
number.
The problems of calculating a correct dimensionless wall distance, y+, in the flows of
drag reducing surfactant solutions exhibiting a complex shear viscosity behaviour were
illustrated by Bewersdorff & Ohlendorf (1988). Fig. 6.6 shows the influence of three
different methods for calculating y+.
When the velocity profile is differentiated to estimate the local shear gradient

(6.3)

where Vs means the solvent viscosity, and the proper shear viscosity of each shear
gradient is taken from a flow curve to calculate the dimensionless wall distance, y+, the
velocity data follow Virk's ultimate profile in the buffer zone. This is a strong indication
that in the turbulent flow of drag reducing surfactant solutions the viscosity can vary
across the cross-section of the pipe.

60 T------.,------,-1 -,----,--:

y
I 4 Re = 120 000 ! . I I /
u ' I
: R;;t~~slooo I
1 ... I
, I
I I
A
I
40 t---t--t-1-!----,
I ... /
I
! !
'
I
vv
I
lv
. I i I
/h
II
B

! I ! i
20

!---,
I
Li~; 1
g
'
.. i"
I I !
i
I , !
! .."-~
.
/ 1'lI
1 ..
I
'

0100
...........,......~... -r- -i I I I I'
2 L 5 B 10 1

Figure 6.7 Dimensionless velocity profiles of an equimolar mixture of hexadecyl-


trimethylammoniumbromide and sodiumsalicylate at c = 830 ppm in a 50 mm
pipe
166 CHAPTER VI

Fig. 6.7 shows two turbulent velocity profiles for the same surfactant system as shown in
Fig. 6.6. In this plot the method described above for calculating y+ was used. At Re =
25'000 the dimensionless velocity profile exhibits only a small buffer layer, whereas at
Re = 120'000 the velocity profile becomes S-shaped, like the velocity profiles close to
maximum drag reduction conditions in Fig. 6.4. This indicates that the shear viscosity in
the buffer layer is mainly influenced by the time averaged velocity gradient and not by the
turbulent fluctuating motion. Attempts have been made to relate the shape of the velocity
profile to different characteristic regions in the friction behaviour, see Bewersdorff &
Ohlendorf (1988).
Apart from velocity profile measurements, additional information is needed in order to
understand the mechanism of the interaction of these additives with turbulence.
Turbulence intensity profiles in drag reducing surfactant solutions were reported by
Povkh et al. (1975, 1980, 1988), Bewersdorff & Ohlendorf (1988), Chara eta!. (1993),
and Beiersdorfer et al. (1994). The maximum of the axial turbulence intensity, which is
found at y+ = 12-13 in flows of Newtonian fluids is shifted with increasing drag
reduction to higher y+-values in surfactant solutions. Depending on the flow regime of
the friction factor versus the Reynolds number curve, both decreases and increases of the
axial turbulence intensity, normalized by the friction velocity, u't, occurred, Povkh et al.
(1980). This behaviour of the axial turbulence intensity js similar to that found in drag
reducing dilute polymer solutions, Berman (1978).

5 surfactant Re- 9600 x Re=l2500 Re=lSOOO


water x Re= 9400 cRe=l2500 c Re=l5000

4 1-

3 1-
.
2 - X
. e9
0
-I(

.Jl'

.
X X 6

.. -
X~- Cf
1 1- ~- X
,..M CIO
.._I(

0 I I

1 10 100 1000

Figure 6.8 Normalized axial turbulence intensity of the same surfactant solution as shown
in Fig. 6.5
DRAG REDUCTION IN SURFACTANT SOLUTIONS 167

Fig. 6.8 shows the behaviour of the nonnalized axial turbulence intensity as a function of
the dimensionless wall distance as found by Beiersdorfer et al. (1994). In the vicinity of
the wall up to y+= 10 the turbulence intensity in surfactant solutions is damped in
comparison to a Newtonian fluid, whereas in the buffer layer an increase of the axial
turbulence intensity occurs.

1.2
surfactant Re= 9600 Re=l5000
water co Re= 9600 .,. Re=15000

Wz 1
ut
0.8 r- c
..,a
0.6 .. c

0.4-
.
0.2 -
c.,.o
... -
... ..;.

0 I I

1 10 100 1000

Figure 6.9 Nonnalized transverse turbulence intensity of the same surfactant solution as
shown in Fig. 6.5

The transverse turbulence intensity, even nonnalized by Ut, is drastically damped in


surfactant solutions in comparison to a Newtonian fluid. It is reduced over the whole
cross-section of the channel or pipe, as shown in Fig. 6.9. Its maximum occurs at the
centre of the pipe or channel, i.e., there is no longer a weak maximum at y+=70, as
found in Newtonian turbulence.
The turbulence intensity measurements show that the vertical (in channels) or radial (in
pipes) momentum transport is reduced in flows or drag reducing surfactant solutions. In
comparison to drag reducing polymer solutions this damping is more pronounced close to
the wall in surfactant solutions. Measurements of the higher moments of turbulence, the
skewness and the flatness, which were done by Povkh et al. (1988) and Beiersdorfer
(1993) in a channel flow of a surfactant solution, demonstrated that the zero crossing of
the skewness and the minimum of the flatness were shifted to higher dimensionless wall
distances in comparison to the Newtonian solvent. Therefore, the zone of the maximum
168 CHAPTER VI

turbulent energy generation also occurs at larger wall distances. Furthermore, it was
found by Povkh et al. ( 1988) that in the zone of the maximum energy generation a large
percentage of the fluctuations are weak and the number of the strong fluctuations is
smaller than in Newtonian turbulence. Integral scales and the microscales of turbulence
were calculated from the autocorrelation function of the axial velocity fluctuations in pipe
flows of drag reducing surfactant solutions by Bewersdorff & Ohlendorf (1988). The
microscale of the surfactant solutions was found to be increased over the whole cross-
section of the pipe in comparison to the Newtonian solvent. The largest increase of the
microscale occurred in the near-wall region. These results can be interpreted as an
increase of the eddies in the dissipation range, which could be caused by an increased
local effective viscosity. The integral scale in the surfactant solution was also increased
over the whole cross-section of the pipe. This means that the large-scale structure is
affected as well, i.e., the structure of turbulence changed at all scales.

0.6 r---------------,
cc +
+

0.4 +
+

0.2 +

0
..

1 4 10 40 100 400 1000
y+

Figure 6.10 The cross-correlation coefficients or the Reynolds shear stressses for the
same surfactant solution as shown in Fig. 6.5, normalized by the product of
the turbulent intensities

The turbulent momentum transfer is characterized by the Reynolds shear stresses. The
distribution of the Reynolds shear stresses, -p u'v', can be calculated in pipe flows of
Newtonian fluids by

u'v' y dU+
--=1----
u.i R dy+
(6.4)
DRAG REDUCTION IN SURFACTANT SOLUTIONS 169

where R is the radius of the pipe. Measurements of the Reynolds shear stresses were
done in pipe and channel flows. Fig. 6.10 presents results from Beiersdorfer et al.
(1994), obtained in a quadratic channel in the form of the cross-correlation coefficient.
The results show a nearly complete damping of the Reynolds shear stresses over the
whole cross-section of the channel, due to a decoupling of u' and v', which is even
stronger than observed in drag reducing polymer solutions.
The results of the Reynolds shear stress measurements in flows of drag reducing
surfactant solutions demonstrate, when the solvent viscosity or the viscosity at low shear
rates is used for calculating the viscous stresses, that a Reynolds stress deficit, G, occurs

u'v' y dU+
--=1-----G
u~ R dy+ (6.5)

As shown by Bewersdorff (1990), this deficit in the momentum balance could be due to
the increase of the shear viscosity in the shear induced state.

2 r'S (5 y"-<0 y'.(Q

0 ~ ti\f ":@ ~
2 -2

f'10 J"10 y'-60 f'60


ttl[SJ qg>~ . .--'
:_~, ~~.

-2 -2

y'a15 r't5 y'.SO y'.SO

ce. -m ~ ~

-
2 -2

2 J"20 y'.ZO J"110 y'a110

-2
tJftF80.2["=i11 ~ 0
2

y'c30 y'~O y'150 r'150

w; . ,.~2J?Nm ~ 0
2 -2
5 -5 -5 -5

Figure 6.11 Joint probability density function for the same surfactant solution as shown
in Fig. 6.5
170 CHAPTER VI

In order to understand the decoupled velocity fluctuations in drag reduced flows of


surfactant solutions, one can study the joint probability density function in these flows in
comparison to that of Newtonian fluids. This was done by Beiersdorfer et al. (1994) and
is shown in Fig. 6.11.
At the location of the highest turbulence production the axial turbulence intensity, u', is,
as already mentioned, much larger than v'. In surfactant solutions the fluctuations become
more anisotropic with the exception of those close to the wall and in the outer flow field.
This plot also provides information on the angle of the main structures in the turbulent
flow with the wall. This angle becomes 180 for surfactant solutions, i.e., the
fluctuations are parallel to the wall.
This result is in agreement with the small-angle neutron scattering results mentioned
before, which showed that at maximum drag reduction conditions the micelles are
oriented in the mean flow direction.

6.3 Rheological properties influencing drag reduction

The flow behaviour of surfactant solutions depends on the temperature as well as on the
surfactant and additional electrolyte concentration. As already shown in Chapter 3.5, a
characteristic rheological feature of drag reducing surfactant solutions is the so-called
shear induced state, an increase in shear viscosity, which occurs above a critical shear
rate.
Since the friction behaviour of drag reducing surfactant solutions is influenced by the
same parameters as the shear viscosity, it is essential to consider the relationship between
the shear viscosity behaviour and the structure of turbulence, which includes the
posssible mechanism of interaction. This includes the elongational viscosity, which
would also be a candidate responsible for the interaction of the additive with the
turbulence. Indeed, the few experimental attempts at measuring the elongational viscosity
of drag reducing surfactant solutions available show an increase in elongational viscosity
in comparison to that of the solvent. However, when the influence of pre-shearing on the
elongational viscosity was studied with the modified rotational viscometer desribed by
Bewersdorff & Vissmann (1990), the Trouton ratio became three when the shear
viscosity in the shear induced state was used to calculate it. Nevertheless, it should be
mentioned that the elongational rates in this study were quite high, i.e., the material
behaviour could be different at low elongational rates.
On the other hand, the measurements of the structure of turbulence in surfactant solutions
at maximum drag reduction conditions reveal a completely changed structure. Burst and
sweep motions are no longer the essential features of the near-wall turbulence. There
DRAG REDUCTION IN SURFACTANT SOLUTIONS 171

like in bursts, occur. Thus, the elongational viscosity of drag reducing surfactant
solutions seems to be of minor importance.
As a consequence, two non-Newtonian material properties remain as candidates for the
interaction of the material with the turbulence: the increased shear viscosity and the first
normal stress difference in the shear induced state. Here it should be mentioned that a
measurable first normal stress difference occurs only in the shear induced state.
Nevertheless, there should also be a non-vanishing first normal stress difference in the
other shear-rate ranges, because stiff rods which become oriented in shear flows cause an
entropy elasticity and, thus, a non-vanishing fust normal stress difference. However, due
to the small number of rod-like micelles in these dilute surfactant solutions, this normal
stress difference should be quite low. Thus, it is not surprising that it cannot be resolved
with the available commercial rheometers.

10 2 ~--------------------------------------~

Nl[Pa] 't(t) [Pa]


A w w X

ioO

10-1

w-2 '-------'-----"----'--'-'--'-'-_w_------"----'--'--'---'-L.L.w 10-2


w-I loO 10 1

Figure 6.12 Rheological behaviour of an equimolar mixture of hexadecyltrimethyl-


ammoniumbromide and sodiumsalicylate at c = 1.8 mmoVl andT = 20 Oc

Fig. 6.12 shows the shear stress and the first normal stress difference of a surfactant
solution which exhibits drag reduction in the turbulent flow regime. These data are
equilibrium data, i.e., in order to exclude time effects, only the steady state values of 't
and N 1 were taken. It should be mentioned that at low shear rates it takes hours to reach
this steady state. Before the equilibrium is reached, 't as well as N 1 exhibit oscillations
during their increase with time. In the steady state the first normal stress difference is
172 CHAPTER VI

during their increase with time. In the steady state the first normal stress difference is
about 200 times larger than the shear stress. Furthermore, it seems that in the shear rate
range studied neither material properties depend on the shear rate. These results show that
the alignment process of the micelles in flow direction causes a large entropy elasticity
which manifests itself in the first normal stress difference. Here, it has to be remembered
that this is due to the superordered structures of the micelles and not due to the individual,
relatively small, single, rod-like micelles.
The increased shear viscosity in the shear induced state can be responsible for the shift of
the dimensionless velocity data to lower dimensionless wall distances, as shown in Figs.
6.6. and 6.7. This way the velocity profiles in the buffer layer become more similar in
shape to Virk's ultimate profile. However, due to the influence of the geometry of the
viscometer on the shear viscosity measurements, which were mentioned and discussed in
Chap. 3.5, the curves of the dimensionless velocity profiles exhibit a certain uncertainty.
Whenever a shear induced state with an increased shear viscosity occurs in a turbulent
flow, the size of the smallest eddies has to increase, and thus, the dissipation rate has to
decrease. Thus, the increased shear viscoity in the shear induced state must influence the
structure of turbulence.

6.4 Hypothesis on the drag reducing mechanism by surfactants

The turbulence structure available, small-angle-neutron-scattering results, and rheological


measurements reveal the following picture at maximum drag reduction conditions: The
micelles become aligned completely in flow direction with their long axis. The turbulent
flow exhibits large fluctuations in flow direction and only small fluctuations in the
direction normal to the wall. These velocity fluctuations are nearly completely decoupled.

6.4.1 A hypothetical rheological explanation

Since the small-angle-neutron-scattering pattern in a turbulent flow at maximum drag


reduction conditions becomes similar to the one in a laminar Couette flow in the shear
induced state, the shear induced structures have also to be built-up in the turbulent drag
reduced flow. In the shear induced state the individual micelles form a new phase of a
superordered structure which exhibits an increased shear viscosity and a large first normal
stress difference, i.e. the surfactant solution becomes viscoelastic.
All published velocity profiles of drag reducing surfactant solutions show a small core
region. This may be due to the fact that the velocity gradients of the mean flow in the core
region are too small to build up shear induced structures. Thus a small turbulent core in
the velocity profiles remains. In the area between the turbulent core and the buffer layer
DRAG REDUCTION IN SURFACTANT SOLUTIONS 173

behaviour. On the other hand, at onset conditions the critical shear gradient for the
occurrence of the shear induced state was found in the buffer zone, which means that the
material seems to behave like a Newtonian fluid in the viscous sublayer. This is also in
agreement with the results of the LDA- measurement of Beiersdorfer et al. (1994), in
which the viscous sublayer was found to reach up to y+= 5. Thus, it is the buffer zone in
which the interaction between the shear induced new viscoelastic phase of the micelles
interacts with turbulence. LDA-measurements and analysis of coherent structures in a
smooth and in a rough boundary layer by Lycko (1993) tend in this direction. The large
velocity fluctuations occurring in the buffer layer could be due to a destruction and
recombination of the "supermicelles" of the new shear induced phase. Thus the origin of
these velocity fluctuations is totally different from that observed in a turbulent boundary
layer of a Newtonian fluid. If one wants to describe the interaction of the new shear
induced phase with turbulence, characteristic time scales of this new phase are needed.
Rheological measurements show that the time scale for the formation and the destruction
of the shear induced state differ by order of magnitude. These time scales of the shear
induced phase have to interact with the spectrum of the structure of turbulence. As a
result, the velocity fluctuations can be suppressed, depending on the time scales of the
shear induced phase. In an extreme case only the largest disturbance can survive. The
large disturbances could be of wavy form and thus lead to large axial and small transverse
velocity fluctuations. This conception by Lumley (1969) is based on an idea by Taylor
(1916, 1960) for a Couette flow. If the viscosity is progressively increased, the extent of
the turbulent region will be reduced, and the viscous sublayer thickness increased. Finally
a state will be reached in which only one disturbance is sufficiently efficient at extracting
energy from the mean motion to stay alive, Fig. 6.13.

u u

-u -u

Figure 6.13 Schematic of Taylor's conceptual turbulent Couette flow experiment, serving
to determine the viscous sublayer thickness, from Lumley (1973)
174 CHAPTER VI

In his considerations Lumley (1969) assumed that the viscosity is progressively


increased, but it did not vary with y+. If the deficit in the momentum balance in the flow
of drag reducing polymer solutions, the so-called Reynolds stress deficit, eq. (5.9), is
explained by an increased local effective viscosity, analysis of available Reynolds stress
measurements show that this effective viscosity increases with the wall distance in the
buffer layer. Berman & Bewersdorff (1988) introduced these calculated effective
viscosities in Lumley's model. The result showed that starting from the end of the
viscous sublayer in a large part of the buffer layer only a small wave number bandwidth
for the turbulence remained, i.e. there were no small eddies at all and the large eddies
were not changed by the additives. It is tentative to hypothesize that for surfactant
solutions the same mechanism can be postulated, especially since a minimum occurs at
y+= 15 where the correlation between these two fluctuation components, u' and v',
vanishes completely.

6.4.2 A fibre-like explanation

The experimental results show that it is difficult to introduce a conception for the
description of the effect which is based on time scales because the occurrence of the effect
is accompanied by an alignment of the micelles. In this respect the micellar systems
exhibit similarities with drag reducing fibre suspensions. In turbulent flows of these fibre
suspensions the alignment of the fibres in flow direction also plays an important role. In
these flows of fibre suspensions a plug flow occurs. Drag reducing polymers and
surfactant solutions exhibit a "quasi-plug flow", in which the fluctuations become
extremely anisotropic and the additives are nearly aligned in flow direction. However,
velocity fluctuation still exist at this maximum drag reduction condition. The alignment of
the micelles must be stable, which means that a disturbance of the alignment has to be
self-correcting. The break-down of drag reduction, which occurs when a critical upper
wall shear stress is exceeded, would be an instability process in this respect, which is
probably similar to the destruction of the plug flow of fibre suspensions.
Since aligned particles can only weakly damp disturbances in flow direction, but strongly
in the transverse direction, this model concept also leads to wavy disturbances. Theories
for such a description do not exist. Attempts were made at the flow of the fibre
suspensions to which polymers were added. Thus the wavy disturbance should be a
significant picture of an asymptotic behaviour. Nevertheless, the goal remains to give the
best description of the interaction of the new shear induced phase with turbulence.
VII DRAG REDUCTION IN FIBRE- AND NON-FIBROUS
SUSPENSIONS

In the flow of paper pulp drag reduction was first observed six decades ago and in fine
sand suspensions three decades ago. The drag reduction in fibre suspensions will be
discussed with respect to fibre dimensions, fibre concentrations, fibre flexibility and fibre
distribution in the flow. As for the other additives, their interaction with the structures of
the turbulent flow, e.g., the drag reduction in fine sand suspensions, will be discussed in
the frame of an inertial theory which shows that drag reduction can occur in a very limited
range of concentration and size. This type of drag reduction also occurs in air flows.

7.1 Drag reduction by non-fibrous suspensions

Non-fibrous suspensions cover suspensions of solid particles of low aspect ratio (length
to diameter). Drag reduction by such particles was observed mainly in gas-dust systems.
A summary of this effect and its explanation was given by Pfeffer & Kane (1974), and
Kane (1990), respectively. A liquid sand drag reducing system was reported by Elata &
Ippen (1961) and Hino (1963). For the effect the relationship between particle size and
the turbulent parameter is important as well as the particle distribution over the cross-
section and their loading.
In liquids the drag reducing dilute mixtures in which the solid loading in percent of the
weight of the fluid is rather small remain Newtonian, with a flow characteristic, however,
showing a thickened viscous sublayer. This means that the particles affect the turbulent
fluctuations, particularly in the near-wall region. Some theoretical conjectures to explain
the drag reducing effect were made and are listed as follows:
(1) A decrease in viscosity
(2) Electrostatic effects
(3) General suppression of turbulence
(4) Stabilization of the transition from laminar to turbulent, and a stabilization of the
near-wall structures of a turbulent flow, respectively
(5) By introducing new length and time scales for the turbulent characteristics
(6) Local turbulence effects resulting in a thickening of the viscous sublayer
(7) A reduction of the near-wall particle concentration by the Magnus effect on the
particle. A lift force on the particle due to the suppressed vorticity by the partic~es.
This rather lengthy list is given to show how many ideas some even contradictory have
been developed to explain the drag reducing effect by non-fibrous particles.
175
176 CHAPTER VII

The first two conjectures were postulated with respect to gas-particle interaction only.
They both contradict the physics involved. The idea behind the first conjecture was that
the particles reduce the free path length of the gas molecules and, therefore, produce a
decrease in viscosity, thus contradicting the Einstein equation (3.37), which predicts an
increase in effective gas viscosity. It has been assumed that by an electrostatic effect
particles could produce a depletion in the particle concentration near the wall and,
subsequently, a decrease in the pressure drop. But this contradicts the physical behaviour
of charged suspensions in a pipe flow. Pipes made of isolator material would have to act
completely different from pipes made of metal. Such an effect has not been reported.
Furthermore, no depletion could be observed, so that conjecture 7 also cannot be the
explanation.
Suppression of turbulence (3) could be achieved if the particles reduced the turbulent
mixing length, especially if the solid particle impact were the reason for most of the wall
shear stress and not the viscous shearing. The idea is that the particles can change the
turbulent structure. It has been assumed that by a general suppression this shows up in
the turbulent fluctuation intensities. Such a result could not be found, see Kane et al.
(1973).
Another variation of the turbulent structures could be a change in their length and time
scale, which results in a thickening of the viscous sublayer. This thickening of the
sublayer was observed and explained by Lumley with dimensional arguments, as
sketched in chapter 5.5.1. He suggested that drag reduction by particles and fibres in
suspension flow was most likely due to an interaction of the particles with the structures
of the buffer layer just above the viscous sublayer. The thickness of the sublayer is, as
shown in chapter 5.5.1, an indication of the size of a disturbance that can be maintained at
equilibrium without input of energy into the sublayer. Smaller disturbances will be
damped by the viscous dissipation. The presence of particles would increase the
dissipation and permit larger disturbances to stay in equilibrium and, hence, permit an
increased sublayer thickness.
The change in the dimensionless sublayer thickness, y5+, when a layer consisting of the
viscous sublayer and the buffer zone is considered, was approximately given by Lumley
(1964) as

(7.1)
DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 177

for very large Reynolds numbers in a turbulent boundary layer with UQ being the bulk
mean velocity.
For example, for Newtonian flows with y 8+,12.5, K"' 0.4 and uQiu,."' 26, the right side
of eq. (7.1) becomes"' -10/13. This shows that doubling the thickness of the sublayer
just about halves the skin friction. Provided the length and time scales introduced by the
particles did not affect the structure outside the sublayer, the von Karman constant, K,

would remain unchanged as well as the logarithmic velocity profile, which would exhibit
a parallel shift to the Newtonian flow only. Thus the main difference compared to the
considerations in chapter 5.5.1 is that not only the viscous sublayer is thickened, but the
buffer layer is displaced outwards away from the wall.
The physical idea behind this interpretation of the drag reducing effect by suspensions is
that the particles have a relaxation time with respect to flow accellerations or
decellerations. The particle react with an inertial effect. The relaxation time is too large to
respond to the smaller energy producing eddies near the wall and will act as an additional
dissipative mechanism. Therefore, panicles can serve to increase the size of the largest
possible equilibrium disturbance in the sublayer and increase the size of the sublayer
itself. Further away from the wall, the size and time scales of the energy producing eddies
increase and the particles may respond to the eddies in this region. It is, therefore,
possible that the same size particles which provide a net increase of dissipation in the
sublayer will also provide a net increase of turbulent production in the turbulent core.
A model which includes the direct interaction of the solid panicles with the flow was
given by Hosaka (1978), who calculated the perturbation of the complete turbulent flow-
field in a gas-solid suspension flow in pipes. The perturbation analysis was based on the
Prandtl mixing length model for the non-interactive flow and incorporated the drag and
the wake behaviour of the particles. With this model Hosaka was able to relate the bulk
mean velocity of the flow to the friction factor. The calculated results compare very well
with experimental data. An analysis of the interaction of particles with decaying vortices
as the main elements of a turbulent flow field was reponed by Gyr (1963), showing that
the relaxation time of the particles is not the only time scale which can change. Gyr
showed that the time scale of the vorticity diffusion process can be relevant, too. A vonex
interacting with suspended particles, for example, decays in a different way. It produces
an area of anisotropic concentration of particles and, therefore, also in viscosity, which
again influences the vorticity diffusion mechanism. Gyr showed that the length scale
changes ( 1967) and also how the time scale changes (1968) due to this kind of
interaction.
Since the inertial as well as the concentration effect act as an increased dissipation at small
vortices, the explanation remains the same for both mechanisms.
178 CHAPTER VII

Theories based on micromechanical interaction processes would be a possibility to


provide the frame work for a more refined modelling of the fluid-solid systems.

7.2 Drag reduction by medium size fibrous suspensions

In regard to their aspect ratio, 1/d, fibrous suspensions were thought to be a kind of
analogical model for dilute rod-like polymer solutions as well as for the micelles in
surfactant solutions; nevertheless the particles are about two orders of magnitude larger.
However, this is only one aspect of the additives and it seems rather improbable that a
complete similarity could be achieved by comparing the three different classes of drag
reducing additives with regard to this parameter only. In a very extensive review Radin et
al. (1975) listed the different influences of solid-fluid systems on the effect. These data
were supplemented with recent data by Kane (1990). However, no basically new ideas
were added.
The main parameters and behaviours relevant for this kind of drag reducing additives are:
1. The aspect ratio, the ratio between the length, 1, and the diameter, d, of the fibre. It
was found that with increasing aspect ratio for the same concentration the drag reduction
increases more pronouncedly, though even more for smaller d. This fact is compatible
with findings that drag reduction increases with the flexibility of the fibres.
2. For a given concentration there exists a Reynolds number for which drag reduction
becomes a maximum. (See also Moyls and Sabersky 1978, who showed that this effect is
even more pronounced in rough pipes.) The reason for the latter fact seems to be that the
stability of the fibre flows is more sensitive to disturbances by the roughnesses than by
the interaction with the turbulent flow over smooth walls.
3. The effect depends on the pipe diameter, D. That reflects the fact that the flow of a
fibre suspension cannot be described by a Reynolds similarity only (non Newtonian
behaviour). Parameters like concentration and fibre type have to be considered too. For a
given fibre type and concentration the onset of the effect is approximately given by an
empirical relation based on ReQ with the pipe diameter and UQ as

)
1
Re0.2
= u~D) = C~v 4
Qo
- - = const.= c (with Re 0 :. uQo
UQ (7.2)

where UQo is the mean flow velocity at onset condition.


These results can be understood if one analyses flow visualizations of the drag reduced
flow, Fig 7.1, together with the appropriate velocity profiles, Fig 7.2, and the friction
DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 179

factor as a function of the Reynolds number, Fig. 7.3. However, the friction factor for
any type of fibre is a function of UQ, D and Cf, the fibre concentration.

Figure 7.1 The visualization of a fibre loaded suspension in a perspex pipe of identical
diameter and with the same concentration, c = 1500 ppm, at three different
Reynolds numbers, (a) 20 000 (b) 40 000 (c) 80 000

Fig. 7.1 shows a visualization of a pipe flow of a suspension of 3F-asbestos fibres in


water. The F-code is a Canadian sieve standard defining the length distribution of the
ISO CHAPTER VII

fibres. 3F stands for a product which belongs to the upper limit of the medium-sized
fibres. Due to their characteristic behaviour with respect to the Reynolds number of the
flow, the flow of such fibre suspensions is classified in three more or less distinct
regions, which also appear in the friction plot, Fig. 7.3:

1. A plug flow (I). A plug of fibres is observed surrounded by a fibre-free annulus of


fluid in laminar flow, Robertson & Mason (1957). In the friction factor plot, Fig. 7.3,
this region is characterized by a straight line with a slope steeper than -1 to the right of
the laminar curve of the solvent and intersecting with the von Karman line. This is
equivalent to a shear thinning behaviour because the Reynolds number is based on the
solvent viscosity. At the same Reo either an increase in concentration or a decrease in
velocity will cause the friction factor to increase. However, it is necessary to check
that the fibres are well dispersed, which means that even if they are entangled, they
have to act as single fibres. It is believed that they are never completely dispersed, as
can be seen in Fig. 2. 7, where a number of fibres is strung together.
2. Mixed flow region (II). The fluid in the annulus was observed to become turbulent,
Robertson & Mason (1957) and Mih & Parker (1967), and the plug begins to
disintegrate in the high shear region at the annulus-plug interface. With increasing UQ
the plug diameter decreases and alsq the friction factor decreases with increasing Re,
however, at a slower rate than in the plug flow region, but faster than in a Newtonian
fluid in a turbulent smooth pipe flow.
3. Fully turbulent region (III). In this region the friction factor no longer decreases with
increasing Reynolds number as rapidly as in a Newtonian fluid, and in many cases the
friction factor becomes nearly constant or even increasing, and the friction curve will,
therefore, reintersect with the von Karman line, hence showing a drag increase. This
result, as Fig 7.3 shows, is not experimentally confirmed. It is an extrapolation of the
available data.

y y y
r---

- u u u

Figure 7.2 The three velocity profiles corresponding to Fig. 7.1


DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 181

The effect is the result of a phase separation and belongs to a class similar to drag
reduction of type B, discussed in chapter 5.5.4, and shows up in the mean velocity
profiles. The logarithmic part of the profile is not shifted, but instead its slope is changed,
in agreement with the observation that this kind of flow most drastically changes the bulk
flow. This was also the reason why Lee & Duffy (1976) for the purpose of
approximating the velocity profile proposed replacing the von Karman constant, K, by an
apparent von Karman constant, Ka, which is a constant for a particular fibre type and
concentration (provided UQ is high> 5m/s).

UQ 1(
- = - I n Re-
u't X:a u't
5.6)
UQ) + (14--
Ka (7.3)

I II III
f

0.010

0.006

0.003

Figure 7.3 Friction factor, f, versus Reynolds number, classifying three regions (I, II,
III) corresponding to the three characteristic flow visualizations shown in Fig.
7.1; _ The von Karman line, and _ _ the laminar friction line,
respectively (From Radin et al., 1975)

Drag reduction is here presented in a phenomenological form. A theory would have to


describe the hydrodynamic interaction based on the orientation of the additives from basic
principles, particularly for semidilute solutions. This is still an unsolved problem (see the
comment in Hinch & Leal 1976). However, recently this problem has been handled for
some simplified flow conditions, such as, for extensional flows, Shaqfeh & Koch
(1990). Shaqfeh & Koch presented a kinetic theory on the dispersion of the orientation of
stiff fibres solely as a result of hydrodynamic interactions about the principal axis of
extension in uniaxial and planar extensions. They found for the mean-squared
displacement of the orientation vector of the fibres to the principal strain axis in the dilute
182 CHAPTER VII

regime, which was defined by eq. (2.1), that it is of the order 0[n0fln2(Lfd)], with n
being the fibre number density. For semidilute suspensions they predict a dispersion of
O[ln(nL3)fnL3]. Thus, the dispersion increases as the concentration is increased from
infinite dilution and then, ultimately, decreases in the semidilute regime.
Since in a turbulent flow local extensional flows of the kind investigated by Shaqfeh &
Koch are always present, their results are a hint at how to understand the setting-up of a
network by the high dispersion of the particles, whereas for higher concentrations the
phenomenon of non-local screening is responsible for the decrease in the oriented
dispersion equivalent to an alignment of the fibres.
In rheological terms the most interesting consequence of the result is the creation of
seco.nd normal stress differences in a planar extensional flow due to the presence of fibre
material in the suspension. Ignoring the contribution of the Newtonian solvent, the ratio
of two viscosity coefficients, 112. and 111, can be considered to describe the rheology of
materials in a pure extension. They are defined as

112 = (0"22 -cr33) I En (7.4a)


111 = (cru- cr22) I Eu = cru I Eu (7 .4b)

using the suspension average velocity field, <ui>, described by a single tensor, Eij

(7.5)

and by taking into account the small dispersion of fibre orientation. The result for dilute
suspensions is

(7.6)

or for semidilute suspensions

(7.7)

In the absence of interparticle hydrodynamic interactions, the fibre material would not
contribute to '112
DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 183

7.3 Drag reduction by long fibrous suspensions

With asbestos fibres it was possible to investigate the influence of fibres with an extreme
aspect ratio, which is for asbestos fibres of the order of 1Q5. In other words, asbestos
fibres are thought to be real macroscopic equivalents to the polymer molecules which
produce high drag reduction, and indeed Ellis (1970) found a drag reduction of 44% by a
concentration by weight of 100 ppm of asbestos fibres. However, he and all his
successors needed surfactants to disperse the fibres, and, therefore, fibre suspensions of
this kind are to a certain extent always a mixture of fibres and surfactants. In addition, by
combining both macroscopic fibres and drag reducing polymers in one suspension, Lee et
al. (1974) found that when both types of additives were used together, they reduced drag
by amounts greater than the sum of the two independent effects. Drag reduction of over
95 % has been reported for such combinations indicating that the two additives may
interact differently with the turbulent flow, although one might think that this is not the
case since the essential qualities for drag reduction are the same for both additives.
McComb & Chan (1985) studied both the influence of fibres on the turbulent flow as well
as the influence of a combination of asbestos fibres and polymers (Separan AP 30). They
found two pronounced regimes of drag reduction: A regime quite similar to the one found
for flows with n{edium-sized fibres, which they call fibre-like, and a flow regime which
has all the characteristics of a drag reducing polymer solution, a so-called polymer-like
regime.
The fibre-like flow regime is present at low Reynolds numbers (1.4104) and in very
"fresh" solutions. In McComb and Chan's experiment this corresponds to the first two
passes of the solution through the system, whereas at higher Reynolds numbers (3.2104
or 5.3104) or higher passes at low Reynolds numbers the flow becomes polymer-like.
The combination of asbestos fibres and polymers shows an unexpected fibre-like
behaviour even for much higher passes through the system. McComb & Chan explained
this behaviour by a rapid degrading of the long fibres due to stress which is inhibited by
the polymers. If stress increased on the fibres with increasing Reynolds number, this
would enhance their degradation and, therefore, also explain their Reynolds number
dependence. This view seems correct if the fibres are more bent since the fibres
themselves resist very high stresses, but not bending. A competing hypothesis would be
a model in which the asbestos fibres are primarily entangled and produce a plug flow, as
do the medium-sized fibres discussed in chapter 7.2. The polymer-like behaviour starts
when the fibres start to act as individual, very long particles. The combination of fibres
and polymers could be explained as an inhibition of a disentanglement of the fibres due to
the polymers which also entangle with the fibres. A hint in this direction is the discovery
184 CHAPTER VII

by McComb & Chan (1985) that drag reduction in their experiments was not the sum of
the drag reduction by both contributors, but practically only by the the fresh fibre
contribution. A much more dynamical interpretation can be given by applying the fmdings
of Harlen & Koch (1993). For a suspension of fibres in a dilute polymer solution under
simple shear conditions at high Deborah numbers, De= y/tc with K the relaxation rate of
the additives andy the shear rate, they showed that the two additives interact to produce a
dramatic increase in the extensions of the polymers, similar to the coil-strech transition
observed in extensional flow. The fibres are perturbed in the angular velocity, giving rise
to a net drift across Jeffrey orbits the orbit of an ellipsoidal particle in a linear flow,
(Jeffrey, 1922) towards the vorticity axis. A model for low Deborah numbers was given
by Leal (1975) for fibres suspended in a second-order fluid. Unlike the results of this
investigation, Harlen & Koch showed that at high De numbers the flow does not depend
on second normal stress difference. If one, therefore, assumes that the polymers are most
active if strained and fibres if properly oriented, the interaction of both types of additives
tends to manipulate the flow in the right direction. But it still remains an open question
whether a plug flow or an accumulated effect is responsible for the drag reducing effect of
the combination of the two additives. We prefer the latter idea based on the velocity
proflles available.
In both regimes the friction versus Reynolds number curve shows in deviation from the
behaviour of the medium-sized fibre solutions in form of a distinct minimum. The
differences between the two regimes show up in the mean velocity profiles as well as in
the properties of the relevant turbulent fluctuations.
In the fibre-like regime the mean velocity profiles can be beyond the Virk's ultimate drag
reduction asymptote. This is the case neither at higher Reynolds numbers nor at higher
passes nor in case of a mixture of asbestos fibres and polymers as shown in Fig. 7 .4.
The turbulent fluctuations show the same picture. In the polymer-like regime the axial
turbulent intensities are enhanced. This is typical for already known drag reducing
polymer solutions which have intensities that are higher than the corresponding
fluctuations in water. However, for degrading fibres the axial turbulent intensities
approach the values of water. In the fibre-like regime the axial turbulent intensities can be
lower than the corresponding values of water, and for the polymer/fibre mixture this is
always the case. The circumferential fluctuations do not significantly deviate from the
values for water. The fibre-like suspensions, however, show intensities which are higher
than the corresponding values of water.
The results show even more pronouncedly that the fibre-like state must have to do with a
bulk effect, probably with a flow very similar to a plug flow. But do the fibres in the
polymer-like regime really act as the polymer does? In other words, at which scales do
the fibres interact with the flow when drag reduction occurs?
DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 185

50

40_

u+ 30_

20_

0
1 100 1000

Figure 7.4 Mean velocity distribution in fibre suspension at Re= 1.4 x1o4: 2 passes
(DR= 76%); o 4 passes (DR= 55%), - - - mixed fibre/polymer suspension
passes 1 to 8 (DR= 5 0-59%); at Re= 3.2 x 1o4: ~ pass 1 (DR= 71 %); _ _
water; _ _ ultimate drag reduction asymptote of the velocity profile (From
McComb & Chan, 1985)

The one-dimensional energy spectra at the center line of the pipe flow for the axial
velocity fluctuations are discussed in regard to this question. The power spectra as a
function of wave numbers, E(kx), is normalized, using the relation

with the transformation of the frequency spectra into the wave number domain, using
Taylor's hypothesis

k=2xf
ii (7.9)

The plot of the typical polymer-like behaviour is shown in Fig. 7.5, whereas the fibre-
like behaviour is given by a plot of a fibre/polymer mixture in Fig. 7 .6.
186 CHAP1ER VII

-2
EX 10

-3
10

-4
10

10 1
J<xD/2

Figure 7.5 One-dimensional energy spectra in fibre suspension at Re= 3.2 x 104 at the
center line; ___ water; _ _ fibre passes 1-3 (DR= 71-58%) (From McComb
& Chan 1985)

It can be observed that in the polymer-like case the inertial range of the power spectrum
contains less energy than in the Newtonian case. At high wave-numbers the spectra are
close or even above those in the Newtonian case, and the fibre-like flows also show a
similar behaviour, though with a difference in degree rather than kind. This result shows
that the polymer-like behaviour of fibre suspension differs from that of a dilute polymer
solution, which is characterized by an energy deficit in the high wave-number range.
The reduced level at intermediate wave-numbers is due to a resonant energy absorption by
the fibres at wave-numbers near the inverse of the fibre length. Polymer/fibre mixtures
would, in this model, produce longer fibre agglomerates and, therefore, interact at'even
lower wave-numbers. This would indicate that the individual fibres or their agglomerates
interact dynamically with the eddies of the turbulent flow. McComb & Chan (1982) used
this hypothesis to infer an eddy-fibre length scale, lm. defined by the power spectra, for
the individual fibres and to correlate this length scale with the drag reducing effect by the
relation
DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 187

DRxlOO% = ( 0.71og10 1; -2.6 )xlOO%


(7.10)

which appears to hold for all the Reynolds numbers studied. lm is defined by the
arithmetic mean of the two wave-numbers marking the beginning and end of the reduced
region, ( lm= 1/km).

E
X

10 1
kxD/2
Figure 7.6 One-dimensional energy spectra in mixed fibre/polymer suspensions at Re=
1.4 x 1o4 at the center line; ___ water; _ _ mixture passes 2-4 (DR= 58-
61 %). (From McComb & Chan 1985)

This result is an argument in favour of ihe degradation hypothesis. The polymer/fibre


mixture would not, however, stabilize the fibres against degradation but, rather, would
stabilize the agglomerates, which are more stable against the degrading forces when held
together by polymer bindings. If this is the case, the asbestos fibres act differently from
the polymer molecules. For the fibres, drag reduction can be due to an interaction of the
individual particles with the flow structures, whereas for the polymers this cannot be the
case.
188 CHAPTER VII

The integral length scale, eq. (4.70), and the time scales found by the autocorrelation, eq.
(4.68), respectively, are in the polymer-like case similar to those found in polymer
solutions. They have, in comparison to water, much larger length scales and persist for
much longer. The burst period increases with drag reduction, as known from dilute
polymer solutions. By contrast, in the fibre-like case the integral length- and time scales
show only a very small increase, and their bursting time increases more slowly with
increasing drag reduction, for the same amount of drag reduction.
Consequently, it is believed that the asbestos fibres are drag reducing by a genuine
mechanism which probably has similarities with the mechanism for dilute polymer
solutions. While similar, they are not the same however. The main similarity is also, in
this case, the limitation by Virk's ultimate drag-reducing asymptote.
The combination of fibres and polymers could, however, also be a genuine kind of drag
reduction of the mixture itself because the combined effect is a non-linear one. A
hypothesis of this kind could be based on Harlen & Koch's (1993) investigations of the
behaviour of a simple shear flow of suspension of fibres and dilute polymer solutions at
high Deborah numbers. This means the shear rates are large compared to the relaxation
rate of the polymer.
It is believed that at high Deborah numbers the addition of small amounts of polymer to a
Newtonian fluid can produce a large increase in the viscosity in an extensional flow.
However, in simple shear flow the increase in viscosity is much smaller. Similarly, the
addition of fibres to a Newtonian fluid produces a smaller increase in the shear viscosity
compared to the increase in extensional viscosity of the suspension.
What happens to the shear viscosity of a mixture of both fibres and polymers? If the
fibres and polymers behave independently, then we would expect only a small change in
the viscosity of the suspension. This is, as Harlen & Koch (1993) showed, not the case.
Instead, a possible interaction between the fibres and the polymers gives rise to a much
larger increase in viscosity. The idea is the following: In an extensional flow the polymer
stretches parallel to the extensional axis so that the velocity difference across the molecule
and, consequently, the stretching force increase as the molecule extends. By contrast, in
shear flow the polymer molecules are stretched in flow direction, perpendicular to the
velocity gradient, and so the velocity difference across the molecule remains small. In a
fibre suspension the polymer no longer experiences a pure simple shear flow, due to
disturbances produced by the fibres. If the effect of these disturbances is to rotate the
polymers away from the alignment in flow direction, the shear flow can stretch them
further on. The result is an increase in shear and extensional viscosity.
Since the orientation of the polymer molecules as well as of the fibres is so important,
some short remarks shall ellucidate the situation in a flowing mixture of both. The
orientation of an isolated fibre in Newtonian fluid follows one of a family of closed
DRAG REDUCTION IN FIBRE- AND NON-FIBROUS SUSPENSIONS 189

curves called Jeffery orbits, Jeffrey (1922), depending on its initial orientation. The
distribution of suspension of fibres between different Jeffery orbits cannot be found from
the motion of an isolated fibre, but depends on secondary effects such as interactions
between fibres. In the case of a weak non-Newtonian fluid, the non-Newtonian stresses
perturb the Jeffrey rotation and cause the fibres to drift between Jeffrey orbits. By
experimental observadons on the motion of the fibres in polymer solutions, Bartram et a!.
(1975) found that, in general, the fibres drift towards an alignment parallel to the vorticity
axis. This is in qualitative agreement with Leal's (1975) calculated perturbation of Jeffery
orbits for fluids with a second-normal-stress difference. Harlen & Koch (1993) then
showed that the drift of the fibres towards the vorticity axis for high Deborah numbers is
much more general and due to a mechanism which is independent of the second-normal-
stress difference.
Based on an elastic dumbbell model for the polymers, Harlen & Koch showed that at
sufficiently high Deborah numbers the combined effect of the fibre velocity disturbances
and the mean shear flow produces a dramatic increase in the extension of the polymers,
similar to the coil-stretch transition observed in extensional flow, where, however, the
extensional rate is larger than the rotational rate for simple shear flows. In the absence of
fibres the simple shear is the special case where the extensional rate and the rotational rate
are equal. The molecules tum with a given frequency. If the duration of stay in the non-
extensional position is long enough compared with the relaxation time, the rotatation is
sufficient to prevent the extensional component of the flow from stretching the polymer.
In other words, if the extension of the polymers is the key parameter for drag reduction
by polymers, the fibres enhance their efficiency dramatically since the polymer also gets
stretched in those areas of the flow field where shear is dominant.
The effect of the polymers on the fibre motion is to align it more to the vorticity axis,
though not completely, so that the feedback system on the polymer is still valid. This
result is supported by the experimental measurements of the orientation distribution by
Gauthier et a!. (1971). In other words, the fibres get aligned in such away that they
enhance the helicity density, eq. (4.25), in the flow, which is thought to reduce the
dissipation in a flow, see chapter 5.5.3.
This discussion would be incomplete without mentioning that surfactant-stabilized
emulsion can also produce respectable drag reduction, see Pal (1993). Many arguments
discussed here can also be used to interpret the emulsion effect. However, we did not
review these drag reducing effects as we omitted the discussion of drag reduction by
micro-bubbles. Both deformable additives are not additives of the kind we had intended
to treat in this book.
190 CHAPTER VII

7.4 Additional effects

Like other drag reducers, fibres also reduce the heat transfer between the turbulent fluid
and the pipe walls in rough and smooth tubes. The conditions under which the reduction
in heat transfer or in momentum transport can be observed are limited. These conditions
are usually reduced to a one-parameter dependence, for example the Reynolds number
dependence, as shown for the friction factor in Fig. 7 .3. However, even for a given
concentration, the Reynolds number is not the only parameter determining the flow.
Usually, as an additional parameter, the wall shear velocity divided by a characteristic of
the suspended medium, as Moyls and Sabersky (1978) postulate, is introduced. This is
important if one needs to apply the Reynolds plot for the design of an industrial
application.
For example, it is a fact that twice the ratio of the heat transfer coefficient to the friction
coefficient for a Prandtl number which goes to unity also becomes one for Newtonian
fluid, whereas for asbestos fibres it tends towards much lower values, such as 0.5 or
0.6. This is an indication that the Reynolds analogy, even for Pr = 1.0, is insufficient for
suspended fibres. For suspended fibres the transfer of heat and momentum must belong
to quite different mechanisms.
An explanation could be that the fibres and the fibre network they produce can transfer
forces but not heat.
VIII APPLICATIONS

8.1 Introductory remarks

As soon as the first experiments in drag reduction by addition of small amounts of


polymers, surfactants, and fibres were reported, several proposals were made for the
possible applications of the effect. However, it is a long way from the idea of a possible
practical application to a realization in a running installation. Several restrictions must be
considered. Firstly, there has to be an ecomomic benefit besides the technical one. Thus it
is essential to consider the following points:
(i) One has to spend money on the additives. Even when the additive concentration
needed to produce the desired drag reduction is very low, the consumption of
additives can be considerable in case of high flow rates. In an economic study the
costs for the additives are part of the variable costs.
Besides, one has to think of how to put the additives into the flow. This requires
injection and/or mixing devices. The costs for these devices are part of the fixed
costs. If pumps or stirrers are needed, they will consume energy, which again
increases the variable costs.
(ii) Any additive will harm our environment to some extent. Thus any possible
application requires a study of the risks for the environment.

8.1.1 General economic considerations

As already mentioned, although minute quantities of additives are sufficient to produce a


drastic drag reduction, the costs for the additives can be considerable. In order to get a
better idea of the amounts of additives neeeded one should remember that the
concentration, ppm, means gjm3. Thus in one through-system, like crude-oil pipelines or
sewers, the additive consumption can produce enormous costs. On the other hand, if the
application is in a closed-loop system, additives are needed which are stable against
mechanical and/or thermal degradation. These requirements for suitable additives also
result in their higher price.
Most of the suitable additives are chemically active substances, i.e. they interact with their
environment. Sometimes the use of additives can create additional technical problems.
For example, anionic surfactants produce foams. Consequently, the use of such additives
implies additional costs for the design and maintenance of the system. If the use of
additives is planned, these costs also have to be taken into account, as well as additional
costs for the operation of a dosing system for the additives and the cost for the effect of
191
192 CHAPTER VIII

the additives on the total period of operation of the installation. Finally, the costs for the
disposal of the additives should be kept in mind.
The additives can be added in form of a powder, an emulsion or a stock solution. Since
drag reduction strongly depends on the additive concentration, a dosing system is
essential. Furthermore, the powder, emulsion, or stock solution have to be stored. This
requires space, which can lead to severe problems, e.g. on platforms for off-shore oil
production.
The criteria provided above can be used to do rough calculations of the ecomomic
benefits of the types of application discussed below.

8.1.2 Environmental considerations

The main criterion for environmental considerations is the poisonous degree of the
additives in general. Although in principle even extremely toxic products can be used in
closed circuits, provided appropriate safety precautions are taken, one has to calculate the
economic benefits of the use of the additives and the additional costs for safety. In some
cases, e.g. for very efficient asbestos fibres (see Chapter 7), a practical application is
nearly impossible, because these fibres produce cancer in the lungs of human beings and
animals. This example elucidates that there are risks for the persons working with the
practical application. This requires additional costs for safety measures. Problems also
arise when the fluid in the closed circuit has to be exchanged. One should remember the
enormous amount of money spent on getting rid of the asbestos fibres in buildings. Thus
investigations of the costs and the risks of a disposal of the additives are essential. Here
as well as in the practical application it has to be considered that the additives can react
chemically. This requires additional studies of the poisonous degree of the side- and end
- products of possible chemical reactions. For example, although many polymers are
non-toxic, some of their monomers can produce cancer.
Last but not least, a special aspect of drag reduction by additives should be mentioned:
The use of additives is attractive because only very few amounts are needed, i.e. drag
reduction occurs in dilute solutions. On the other hand, this can create severe problems, if
an additive recovery is desired. This would require the development of a completely new
separation process.
APPLICATIONS 193

8.1.3 Optimimization of the selection process for the additives

In the last two subchapters the problems which can arise in the practical application of
drag reduction were shown. The reason for this was not to deter the reader, because the
number of available additives is so high that for each application a suitable additive can be
found. Many factors influence the selection process. Thus it is helpful to start with an
analysis of the system. In many cases this leads to the requirements which have to be
fulfilled by an additive. A suitable additive may be available, a combination of additves
may have to be used, or the additive may have to be developed. These general
considerations elucidate that in comparison to today's state of the art more biologically
degradable additives have to be developed in order to increase the use of additives in
practical applications.
These warnings or, better, careful admonitions were mentioned before presenting the
applications in order to avoid "errors" in the understanding of the application of additives,
which could lead to a bad image of the drag reducing additives in general. This way the
additives will be broadly accepted for what they are: energy saving additives.

8.1.4 Systematics

In Chapters 8.2 and 8.3 two one-through applications where polymer additives are
currently used will be described. The largest and best-known application is the use of
polymers in crude-oil pipelines for increasing the flow rate at fixed pumping costs. In this
application polymers are permanently injected. This has led to enormous economic
benefits, although large amounts of polymers were needed. The use of polymers in
sewers to prevent flooding, as a result of an overload during times of heavy rainfalls is a
sporadic application. Furthermore, in this example one sees that different environmental
problems can occur simultaneously.
The addition of surfactants in district heating systems or large air conditioning systems is
an example of a permanent use in a closed-loop system (Chapter 8.4).
The use of additives is not restricted to pipe or open channel flows. Also boundary layers
(Chapter 8.5) or flows with separation (Chapter 8.6) can be influenced by the addition of
additives. The injection of polymers into the boundary layer of hydrofoils can be used to
reduce the drag of ships. The hydrotransport (Chapter 8.7) is also an application in which
mixtures of different additives can be used, especially when the transported solid material
exhibits drag reducing properties too, e.g. paper pulp.
Since drag reduction is accompanied by a change in the structure of turbulence, there are
some cases in which additives can be used, but drag reduction itself is not desired. The
194 CHAPTER VIII

changes in the structure of turbulence can prevent cavitation (Chapter 8.8) or stabilize the
sediment transport, i.e. the interactions with an errodible bed (Chapter 8.9).
The most complex applications are those in which biological aspects are important. In
Chapter 8.10 applications in artificial irrigation systems are discussed, whereas in
Chapter 8.11 the first attempts at medical applications are reviewed.
Besides the applications of real drag reduction there exist other applications in which the
changed material properties, due to the presence of the additives, are used. This holds for
the stability of free jets in fire fighting, for the so-called anti-misting agents to prevent
explosions, etc. These types of application will be mentioned, but not discussed in detail.
Finally, possible future applications will be looked at in Chapter 8.12 and in Chapter
8.13 some suggestions will be made for the preparation of drag reducing solutions.

8.2 Drag reduction in crude-oil pipelines

The first use of drag reduction on a large scale took place in the Trans Alaskan Pipeline in
1979, Burger et al. (1980). Here, an oil-soluble polymer, polymerized from a
straitchained a-olefin monomer, is injected as a concentrated 10 % stock polymer
solution into the pipeline downstream of each pumping station, in order to increase the
flow rate at a constant pressure drop. Due to the effectiveness of the polymer injection
technique there was no need to build two further pumping stations. Since 1979 the
polymers used in this application were improved. This improvement led to an increased
effectiveness of the additive, so that the polymer concentration could be lowered by a
factor of 12 since 1979, Motier & Carrier (1989). This was done by optimizing the
polymerization technique of the polymers in order to increase the molecular weight of the
polymers, which is now in the order of 2.4 - 3 107 g/mol. When the molecular weight
is increased, the same drag reduction occurs at a lower polymer concentration, because
the onset of drag reductioin occurs at a lower wall shear stress (see Chapter 5). However,
a polymer of such high molecular weight is extremely sensitive to mechanical
degradation. Thus, the injection system has to be designed in order to minimize possible
sources of mechanical degradation. Nevertheless, there is a so-called on-line degradation
in each pipeline segment, i.e. the local drag reduction downstream of an injection unit
decreases with the distance from it, which means that even the relatively low stresses
present in this large pipeline of d = 1.25 m are sufficient to degrade this sensitive, high-
molecular-weight polymer. Nowadays the polymer concentration in the Trans Alaskan
Pipeline is in the order of 1 ppm, which results in an increase of the flow rate by 33 %.
In 1989, 25 injection sites for using drag reducing polymers for product pipelines existed
in the U.S.A., Motier & Carrier (1989), and many others worldwide, especially at off-
shore crude-oil pipelines, e.g. in the Bass Strait (Australia), North East China, India
APPLICATIONS 195

(Assam and Bombay off-shore oil), the Middle-East, North-Sea, etc. In general, to avoid
degradation, the polymers are injected downstream of a pumping station. Drag reduction
is mainly used to increase the flow rate, i.e. the capacity of an existing pipeline, which
provides the possibility of increasing the oil production. Furthermore, the use of drag
reducing polymer solutions in these applications has another advantage: Since drag
reduction is accompanied by a reduction of heat exchange, the cooling of the crude-oil
during transportation is also reduced when polymers are added. Consequently, the
temperature of the crude-oil can be kept on a higher level, which, due to the lower
viscosity, results in a higher flow rate at a constant pressure drop.
Drag reduction in crude-oil pipelines is a good example for showing how the optimization
process of an application takes place. First, drag reduction was only used because the
construction of two pumping stations was delayed. After successful trials, which showed
that this method worked and produced ecomomic benefits, the two planned pumping
stations were cancelled and, in a next step, the polymers were improved in their
efficiency. Here it was shown that the use of a more expensive, specially designed
higher-molecular-weight polymer is cheaper than spending more money on larger
amounts of lower-molecular-weight polymers.
In this application the use of polymers causes nearly no environmental problems. The
crude-oil into which the polymer is injected has to be refined later on. During the
distillation processes of the crude-oil the polymers are thermally degraded, i.e. cracked
like the crude-oil. Safety precautions at the injection stations and during the transport of
the stock polymer solutions are the only measures to be taken.

8.3 Drag Reduction in Sewer Flows

The discharge in a sewer exhibits large fluctuations, which are due to variations of the
rainfall in time and space and due to the fluctuating use of water in private households
and in industry. In so-called mixed systems, which contain waste water and rain water,
there is only a low filling level in a sewer, because sewers are designed for the short-time
high flow rates during heavy rainfalls. If during heavy rainfalls an overloading of a sewer
occurs, it results in flooding and causes considerable damage. During a sewer's period of
service the frequency of flooding may reach an unacceptable level, because the flow rate
in the sewer increases during heavy rainfalls as a result of an increased catchment, such
as by a more sealed area. It is then necessary to think of constructing a new sewer
system. An alternative would be a polymer dosing during the short time periods of heavy
rainfall, an economically attractive solution because a new construction is capital
intensive. Polymer dosing will increase the flow rate in the sewer. Besides, it is quickly
196 CHAPTER VIII

applicable. By polymer dosing the construction of a new sewer can be delayed or even
made unnecessary.
In pipe flow, for a constant discharge, Q, the drag reduction is defined in eq. (1.6). In
sewers the friction loss is determined by the bed slope of the sewer. Drag reduction leads
to an increase in the bulk velocity. Under full flow conditions this leads to an increase in
discharge. Under free surface conditions with constant discharge the cross-sectional area
of the fluid decreases.
Although the capacity of a sewer is its most important factor, the increase in bulk
velocity, ~UQ. is the best description of drag reduction under full flow and free surface
flow conditions

u -u u
AuQ = P s = __f. -1
Us Us (8.1)

The first polymer dosing experiments in sewers were done in 1969 in Texas by the
Western Company (1969). In their report special attention was given to the interaction of
the dissolved polymers and the biological microorganisms present in a sewer system. Up
to polymer concentrations of 500 ppm (polyacrylamide and polyethyleneoxide) no
damages of the microorganisms were detected. This is in agreement with the results by
Wade (1972), who found that both polymers are chemically and biologically degradable.
Furthermore, the polymer concentration in the sewage-treatment plant will be much less
than in the sewer where the polymer dosing is applied, because polymer dosing will
operate only in a small percentage of the sewers. The flow rate increase in the
experiments described by the Western Company (1969) was between 10% and 30 %.
Based on these successful tests in a sewer with an automatically operating polymer
injection, injection systems were installed by Chandler & Lewis (1977) and by Sellin
( 1977 and 1980). In these practical studies it was found, that with the exception of a few
cases, it was possible to waive a pre-dissolving of the polymer. It was sufficient to mix
the polymer powder in a vortex chamber with water before injecting this mixture into the
sewer. The polymer dosing was started and stopped when a certain filling level in the
sewer was reached. Furthermore, Sellin (1977) showed that often the use of polymer
dosing is cheaper than constructing a new sewer with a larger diameter, a result which
was confirmed by Bewersdorff et al. (1986) by a cost analysis of sewers recently
constructed in the city of Dortmund in Germany. At the Knowle Sewer in Bristol,
England, Sellin (1980) was able to achieve a flow rate increase by 40 % at polymer
concentrations of less than 100 ppm. Sellin (1980) as well as Bewersdorff et al. (1986)
also ran experiments when the sewer was partly filled. When the filling level in a
APPLICATIONS 197

manhole downstream of the polymer injection system was recorded, a typical behaviour
of the level curve was found, as shown in Fig. 8.1.

120~r--------,----------~

h[mm)

100

80

7 t [min) 9

Figure 8.1 Level curve with polymer dosing

20
...
Ql
..s!
%

~ 15 r-
...:s
~

"'
.5
10-
~

..."'
=
...
~

Col
5
.5
0 I I
0 20 40 60 80 100 120
c [ppm]

Figure 8.2Increase in water surface level in a manhole versus the polymer concentration

The first part shows a constant level for steady flow conditions. Then a peak followed by
a new con.stant level can be observed. This wave results from starting the polymer
dosing. The water containig the polymers is accelerated and mixed with the downstream
water, which does not contain polymers. This mixing zone produces the wave. The
following constant level shows the drag reduced steady flow. (In this experiment the
discharge is not constant, therefore the new level cannot be expected to be lower than the
first level of steady flow.) This is followed by a negative wave before the first constant
level is reached again. This negative wave occurs when the polymer dosing is stopped.
The water without polymers is too slow to follow the accelerated polymer-containing
water. This can be seen as the short period with a low level.
198 CHAPTER VIII

The increase in surface level depends on the polymer concentration, as shown in Fig.
8.2, as well as on the solution or mixing time of the polymer powder/gel-water mixture,
as shown in Fig. 8.3.

100
...
Q)
~
%
~
.;::
s..
:::1 so~
.5"'
~

~"'
~
s..
.5"' 0 .... "'-
0 10 25 75 195 t [s]

Figure 8.3 Increase in water surface level in a manhole versus the solution time for
a 60 ppm polymer concentration

40
%
t
10 s .
30 t- +
25 s *
**
~

"' * *

~
~
75 s
s..
"'
.5 * 105 s
..... 20 *
* *
r-
o..... +
.9
... +
. .
~

*
.
10
*#j + +
. ..
0
0 20 40 60 80 100 120
c [ppm]

Figure 8.4 Velocity increase for different polymer concentrations and solution
times

According to Sellin (1978) the filling level in the manholes of the sewer will increase with
the distance from the polymer dosing system. In order to minimize this effect, which
could also cause flooding, Sellin (1978) proposes to start with the polymer injection early
enough, i.e. before the sewer is overloaded, and to slowly increase the polymer
concentration. The behaviour of the sewer level, when polymer dosing is started, was
studied in a numerical simulation by Oles (1989).
The increase in velocity obtained by polymer dosing depends on the polymer con-
centration and on the solution time, as shown in Fig. 8.4. In this figure there is no
APPLICATIONS 199

significant difference between the curves with a solution time of 195s and 75s. Both
increase up to 70 ppm. For polymer concentrations above 70 ppm there is a constant level
at about 30% increase in the velocity. The curves for 25s and lOs solution times show
less drag reducing effects, but exhibit a similar shape. For concentrations above 70 ppm
an asymptotic increase is reached. The maximum increase in the velocity is 17 % for 25s
and 8 % for lOs solution time.
These experimental results demonstrate that the increase of the velocity in the sewer
depends on the polymer concentration and the solution time. From the experimental
results showing the influence of the solution time it may be concluded that a minimum
solution time is necessary to obtain the drag reducing effect. For practical use there has to
be a minimum distance between the point where the polymer is added and the point where
the increase in flow rate is necessary. This can easily be done.

160000
slope
DM/year

120000

80000

40000

0
0 600 800 1000
d

Figure 8.5 Comparison of costs for polymer dosing and a new sewer construction

In order to decide whether the construction of a new sewer or the use of the polymer
injection technique is more economic a comparison of the cost of the two alternatives is
necessary. The costs of the polymer injection technique are mainly caused by the costs of
the polymer, whereas the costs for the polymer injection system are low in comparison.
The annual polymer costs depend on the sewer diameter, the slope of the sewer, the
surface roughness, the increase in discharge, the polymer concentration, and the price of
200 CHAPTER VIII

the polymer. Two of these parameters (surface roughness and price of the polymer) are
usually known. The other parameters depend on the application site.
The costs for the construction of a new sewer are mainly caused by earthworks.
Depending on the nature of the soil these cost can vary to a large degree. In order to com-
pare the costs for the construction of a new sewer with the costs for the polymer injection
technique, annual figures are needed, which require exact information on the sewer's
period of service.
In Fig. 8.5 the annual costs for the construction of a new sewer (buried 3 m, length 1
km), which are inside the section line area, depending on the composition of the ground,
are compared with those of a polymer dosing system (dosing 20 h/year, polymer concen-
tration 20 ppm, polymer price 10 DM.-/kg) for different slopes and diameters of the
sewer. It follows from Fig. 8.5 that the polymer dosing can be very cost-effective,
especially when the slope of the sewer is low and the diameter small. With increasing
length of the part of the sewer which needs to be replaced the polymer dosing technique
also becomes more cost-effective.

8.4 Central Heating Systems

In primary flow circuits of large heating systems, in cooling flow circuits in the chemical
industry, and in air-conditioning systems of large buildings the pumping energy could be
reduced by adding drag reducing agents. This would mean significantly saving energy,
because the flow rates in these installations are usually high. On the other hand, there is a
restriction on these applications: the drag reduction is accompanied by a heat transfer
reduction. In a heating body this does not cause any problems because the largest heat
transfer resistance is at the interface solid - air and not at the interface liquid - air.
However, if it is intended to apply drag reduction by additives in an existing district
heating system consitsting of a primary and a secondary circuit, this can lead to severe
problems. The heat transfer in the heat exchanger may become insufficient. Hence, the
influence of the additives on the heat transfer has to be studied as well as the influence on
the friction.
The hydrodynamic conditions in these installations are quite different from those of other
applications. Normally these completely closed systems operate for periods of months or
even years before the circulating water is exchanged. The diameters of the pipes are larger
than in other applications, the Reynolds numbers are in the region of lOS, and the
temperatures in the circulating systems vary between 5 Oc and 130 oc, depending on
whether it is a district heating or an air-conditioning system. Since the water in these
installations is circled for long periods of time, corrosion inhibitors are normally added.
Thus one also has to check the interaction between these substances and the drag
APPLICATIONS 201

reducing additives. If one intends using surfactants as drag reducers, the influence of the
salts which dissolved in the water of the flow circuit on micelle formations also has to be
considered.
First, polymers were tested in these applications. Besides the problems of mechanical
degradation, which may become serious due to the large number of passes through a
pump, a thermal degradation occurs. However, experimental results reported by Pollert
(1978) showed that the thermal degradation simulated by boiling the polymer solution for
a certain period had only little adverse effects on the subsequent drag reduction. Attempts
to find additives for stabilizing the polymers under technical conditions were made by
Martischius & Heide (1984). They found that an oxygen scavenger, Na2S03, improved
the thermal stability of the polymer. Nevertheless, it did not prevent degradation.
However, an unexpected synergistic effect was found when the oxygen scavenger and a
chelating agent were used simultaneously. At a temperature of 150 Oc a polyacrylamide
solution exhibited a constant drag reduction, even over a period of 21 days.
Successful experiments in district heating systems were reported by Amborn &
Hagstrand (1977), Elperlin eta!. (1971), Leca & Leca (1984), and Pollert eta!. (1982 a,
b). All these experiments report various degrees of drag reduction over long time periods.
Besides, polymer additives surfactants can be used for this type of application (The
research on surfactant drag reduction was stimulated by this type of application). The first
application was reported by Povkh et al. ( 1980), who used a surfactant called
"Methaupon". Since the drag reducing properties of surfactant solutions depend on
temperature, it is very difficult to predict the drag reduction by adding surfactants,
because the temperature varies within these installations. One advantage of using
surfactants is that by changing the chemical components of the surfactant system, the
surfactant system can be optimized for the desired temperature range. For example, when
the length of the hydrocarbon chain of a given surfactant system is increased by adding
alkyl groups, the new system will exhibit drag reduction at a higher temperature level.
Furthermore, by using an appropriate surfactant system dependent on the design of the
heat exchanger the reduction of the heat transfer can be avoided.
As already mentioned, drag reduction in surfactant systems vanishes when a critical wall
shear stress is exceeded. Thus, if the wall shear stress in the heat exchanger is above the
critical wall shear stress and the wall shear stress in the pipes is below the critical wall
shear stress, one may have drag reduction in the pipes and a heat transfer in the heat
exchanger which is hardly or not affected by the additives.
Rose et al. (1984) reported results on an application in a building hydronic heating
system in which the surfactant solution was shown to be stable for two and a half months
of continuous pumping and repeated temperature cycles.
202 CHAPTER VIII

The first successful applications of the use of drag reducing surfactant additives in district
heating systems stimulated research in the chemical industry. Surfactant systems were
found which exhibited drag reduction even at surfactant concentrations of less than 100
ppm. This is documented by several European and U.S. patents. Ohlendorf (1985)
added surfactants to a small district heating system in Aachen/Germany and achieved an
overall drag reduction of 57 % in this system. The experiment lasted one year. There was
only a slight reduction of the surfactant concentration due to adsorption of the surfactant
at inner walls of the heating bodies of the houses. A large research project on "Drag
reducing additives" was sponsored by the German Ministry of Science and Technology.
Investigations of the operational behaviour of several heat exchangers were done, work
on measuring and control experiments, consumer installations, pumps, corrosion and
related water treatment problems were initiated. Research in real district heating systems
was carried out to confirm the effects measured on a laboratory scale by Steiff et al.
(1989) and Fankhanel, M. (1989). These results show that drag reduction in district
heating systems can be applied and that it leads to large energy savings. The use of
surfactant additives for reducing the pumping costs in large air-conditioning systems is
currently under investigation.

8.5 Ships and submerged bodies

An especially efficient method to produce drag reduction for a body in motion in a liquid
is to add drag reducing particles into its turbulent boundary layer. This has been done for
ship hulls and for submerged bodies such as submarines, torpedoes and water-bourne
rockets. Even swimmers use this technique.
Two main methods of feeding the turbulent boundary layer with additives were tested:
The feeding by additive-containing paints or by injection of the additives, mainly
polymers, through slits and nozzles. The second method proved too expensive for
civilian purposes, however, it was used for military applications, i.e., the flee velocity of
a submarine could be instantaneously increased, and the noise reduced, see chap. 8.8.
For civilian applications ship paints were used which were composed of a porous
residual, containing drag reducing polymers in its cavities. The polymers diffusing into
the turbulent boundary layer produced a sizable drag reduction of the ship hull. The ships
covered with such paint were more rapid and consumed less energy compared with a
conventionally painted ship. The problem was that the ships needed a new coat of paint
after a rather short time, i.e., after about two crossings of the Atlantic. The painting of a
ship hull reqires the very costly residence of the ship in a dry dock and therefore this
application failed, too. The situation for small boats and racing boatsis somewhat better.
However, the racing laws forbid paints of this kind.
APPLICATIONS 203

The trend for reducing the drag of ships or racing boats is therefore to use passive drag
reducers, which do not need the same maintenance, such as sheets of riblets, which are
used to cover the hull or part of it, see i.e. Choi (1990). Micro bubbles, a kind of
vanishing additives, are active drag reducers. Lumley (1985) reported on two Soviet
reports. For some Western studies see, e.g. Brown (1990) or Legner (1990).
It is more promising , however, to optimize the shape of the hull. To do this one has to
test ship models in a towing tank. For reasons of similarity these studies are limited in
size. By using drag reducing additives in the tank the Ut can be reduced with respect to
water for the same towing velocity. This allows to test ship models corresponding to a
larger size or to study the velocity behaviour at higher values in the existing tanks.
Injections during short periods is another promising application. Ships normally move on
only half the power of their engines when they are at sea, in order to reduce their power
consumption. Full power is only required for their manoeuvering in harbors or at sea,
when a large number of high beds, rocks or sand banks exist. Thus, if the use of drag
reducing additives is considered during the construction of the ship, the power of the
engine, which has a large influence on the costs, could be reduced.

8.6 Separation and Hydrofoil.s

The hydrofoil craft are a special class of boats. The surface ratio between the hull and the
foil is extremely large, which means that the air is the main drag. However, the drag of
the hydrofoil is important too and therefore attempts have been made to reduce the drag
by additives. This application failed for several reasons. First of all, the hydrofoil is
mainly used in the supersonic mode and it is not yet known how the additives work
under this condition. Secondly, for the lift of the hydrofoils the essential condition is the
kind of separation which the flow around them exhibits. Additives can drastically
influence the location of the separation. Therefore, in extreme cases additives can destroy
the lift completely. This problem is not new. Early experiments with additives injected at
the bow of a ship reduced the drag, but also decreased the efficiency of the propulsion by
the ship propeller. Again, this is due to a change in the separation at the blades.
Nowadays, sensors and computers are available which would allow a controlled injection
of the additives, which takes care of the interaction with the separation. But this has not
yet been done. Some additional remarks will bemade in Chap. 8.12.2.7.
204 CHAPTER VIII

8.7 Hydrotransport

The hydraulic transport of solids in pipes is economically attractive in comparison to


transport by truck, railway or ship. The goods transported over long distances are mainly
coal and iron ore and the carrier fluid is water. Since the longest pipelines for the
hydraulic transport are about 400 km in length, by using drag reducing additives
considerable energy saving should be possible in the hydraulic transport of solids.
The first large scale experiments of this type were reported by Pollen (1977, 1985).
Pollen injected polymers in a pipeline used for the transport of fly ash [rom a thermal
power station to the disposal. This pipeline was 4.1 km in length and 250 mm in
diameter. The volume concentration of the solid was 4 %. By adding polymers the bulk
velocity in the pipeline and the solid concentration of the hydrotransport could be
increased. So this method can be used to transport bigger amounts of fly ash during the
peak values of the thermal power station. Pollen (1978) also reported on the application
of drag reduction to the hydraulic transport of sludges from the central sewage treatment
plant in Prague to a sedimentation field. He added polyacrylamides and found that the
cationic polyacrylamides increase the sedimentation, whereas anionic polyacrylamides
exhibit better drag reducing properties. He concluded that the combination of the two
should be technically and economically more advantageous.
Golda (1986) studied the influence of polymer addition on the hydraulic transport of coal.
He found that the addition of polymers to solid-liquid mixtures also produces drag
reduction. At a constant polymer concentration the drag reduction decreased with the
solid concentration. In coal-water mixtures higher concentrations were needed to produce
the same drag reduction as in a pure water flow. This may be due to the adsorption of the
polyelectrolytes on the surfaces of the coal, which causes a decrease in the actual polymer
concentration in the fluid phase. Due to degradation, a decrease of the drag reduction
occurred with the transport time. It has to be considered that these studies were done in a
closed circuit with a high number of passes through a pump. Thus, the degradation in a
real pipeline should be of less influence.
An application of drag reduction to an existing transport system would require the
knowledge of the influence of the additives on another characteristic property of a
hydrotransport system, the settling velocity of the solid particles. Normally,
hydrotransport systems have to be operated at velocities which are a bit above the settling
velocity in order to minimize the pressure drop. Thus, it is important to know in which
way the settling velocity, in the hydrotranport is affected by the addition of the drag
reducing agents.
APPLICATIONS 205

8.8 Cavitation protection and noise suppression

If in a fluid the local pressure declines below the vapor pressure, small bubbles of vapor
are created. This can be the case in areas of high acceleration in a turbulent flow. If the
flow is afterwards decelerated, and the local pressure therefore increases, the vapor
condensates. This process can be very rapid, and due to the resulting drastic change in
volume the bubbles will collapse in form of an implosion, giving rise to enormous
pressure shocks. Up to 10'000 bar have been measured.
Close to the wall such local underpressures are sources of destruction to the wall
material. Damages by cavitation can therefore mainly be observed on structures or
elements exposed to a flow with high acceleration in its boundary layer. This is the case
on the blades of water turbines, ship propellers, valves, etc.
If cavitation occurs as the result of local events due to turbulence, one can expect that
additives help to reduce or to suppress the bubble formation, since the additives lower the
strain in the critical flow events. This interaction should reduce the size of the bubbles
and increase their wall distance, since the sublayer is thickened in case of a drag reducing
flow. The damages should therefore decrease in comparison to Newtonian fluid, since
the pressure wave at the wall decreases due to the two changes mentioned.
In earlier years several papers appeared confirming drag reducers as protectors against
cavitation. They are listed in White & Hemmings (1976). Sporadic publications on this
theme appeared, without giving deeper insight into the problem though. There are some
practical reasons why this application did not appear very attractive. In most cases it is
not economic to seed the whole flow field around the element. If one wishes to protect a
rotative system, for example, it is necessary to feed the additives onto the blades very
locally, a task which is very complicated. In addition, the effect of suppressing cavitation
can be accompanied by a reduction of the efficiency of the turbines, for example, since
the flow conditions around a blade change too, often due to an undesirable change in
separation, (see also chapter 8.6).
Cavitation is one of the main sources of noise. Therefore, whenever the suppression of
noise is an essential problem, like in many military applications, it is advisable to
consider whether the use of drag reducing additives could be a solution.
206 CHAPTER VIII

8.9 Stabilization of sediment transport

Sediment transport is a geophysical phenomenon which causes problems when it is either


not efficient enough or when it is so efficient that erosion is the result. Rivers of the first
kind are mainly rivers with low slopes and fairly fine sediments, as, for example, the
Yellow River in China. The other type of rivers, also in an unstable condition, mainly
erode because the sediment supply is not in equilibrium with the transport. Many
European rivers became this type because the sediment is retained in one or the other of
mankind's fluvial constructions.
To manipulate rivers by additives would be very attractive, but it is out of the question,
since the amount of material required would be outrageous economic-wise as well as in
respect to the contamination produced. However, the problem would not be an academic
one, only if some suspensions in the river itself were drag reducing. And, in fact, fine
clay suspensions are of this kind, and silt has been reported to produce an elastic layer,
Gust (1976) or Nihoul (1977).
The mechanism is still debated because it seems that it is an effect which fibres and
flocculates produce in combination with electrical forces at the edge of the particles.
Nevertheless, from the overall knowledge of drag reduction it can be argued that for the
same discharge a drag reducing river would have a lower wall shear stress, and therefore
the suspensions would create a hydrodynamical stabilization of the sediment transport.
Gyr & Schmidt (1989) investigated the stabilization of erodible beds by dilute polymer
solutions in an experimental flume for hydraulically smooth bed conditions. To achieve
drag reduction the wall shear stress had to be so high that sand grain would have gone
into suspension. Therefore the sediment material was a lead granulate. The rather
dramatic result was that the bed remained stable even at shear stresses at which the same
sediment was strongly transported in pure water. The explanation was that the sweeps,
the most important events for the sediment transport, are mainly changed in their
direction, see Fig. 5.10. The momentum exchange between the bed and the sweeps
becomes very low because of the flat angle at which the sweeps are oriented towards the
wall.
This result helps to understand the stabilization of the transport in rivers with the drag
reducing suspensions mentioned. On a practical scale this result is important if polymers
are used in sewers, as discussed in chapter 8.3, since practically every sewer channel has
a fine sand bed.
APPLICATIONS 207

8.10 Biological applications

Instead of "biological" applications one would rather expect the term "biomedical".
However, as we will see, the use of drag reducers in nature is on a more general basis
and therefore also the applicability of the effect. For example, some fish or plants (i.e.
okra) produce slime which is drag reducing. This slime or derivatives of it could perhaps
be very interesting biomedical products, since such slime should have a higher biological
acceptability by a living body than the artificial additives normally used as drag reducers.
When one thinks of biomedical applications of drag reducers, one immediately associates
their use with enhancing the circulation of the blood. The main goal is to reduce high
blood pressure, due to arteriosclerotic contractions, for example.
Blood itself is not a Newtonian fluid but is only shear thinning along with a slight yield
stress. The reason for this rheology is that the blood cells are tiny elastic bags of
hemoglobin. Nevertheless, a combination of blood and drag reducing additives, like
polyacrylamids, is a drag reducing solution and, what is important in this respect, it also
remains so in pulsed flows. Together with the fact that the turbulent intensity level is
reduced by the additives, their application becomes attractive. The flow would be
enhanced through the constraint and tromps, which are thought to be created by turbulent
separation at arterial valve membranes, and could possibly be suppressed and the danger
of thromboses thus reduced.
Such applications are not yet practicable and the reason is that the blood has to pass the
capillary vessels during one circulation turn. Their diameter is of the size of one blood
cell, in other words, so thin that it could get plugged up by the additives.
The use of the additives in hard-lung machines or in an artificial kidney is another matter.
In both cases the blood circulates outside the body and is therefore not limited by the
capillaries as long as it is possible to remove them when fed back into the natural blood
circulation. In both cases the blood cells are less stressed and are therefore less damaged,
a serious problem in such treatment. For the same reason it was also proposed to use
drag reducers against atherogenesis. A sort of review of the applications of drag reducers
in the blood circulation can be found in Greene eta!. (1974). See also White & Hemming
(1976).
In short, there is no doubt that drag reducers would be highly beneficial in many medical
applications. However, the appropriate additives are missing. It was mentioned at the
beginning of this chapter that new additives have to be developed. So far only two
suggestions have been made in this respect. One is to look for a slime-like drug and the
second is to produce DNA molecules, of which it is known that they are drag reducing if
208 CHAPTER VIII

suspended in monodisperse form. They can be produced by extracting them from the
patient's own cells.
An unexpected biological application follows from the fact that certain algae are drag
reducing. If such algae could, for example, be economically grown in an upper water
storage basin, these algae could be used as drag reducers in the electric power production
or in irrigation systems, where they could additionally be used as fertilizer.

8.11 Irrigation systems

In many parts of the world artificial irrigation is used in agriculture in order to increase
the production per area. If polymers are found which have no negative effects on the
plants and on the soil, the use of drag reducing agents in these systems, which mainly
use centrifugal pumps, would increase the flow-rate of the output and consequently the
area which is irrigated.
In the sudies of Elias & Vocel (1978, 1980 a, b) polyacrylamides were tested in
Czechoslovakia for this purpose. No negative effect of the additives on the plants were
found. This was confirmed by extensive studies by Singh et al. (1979) and Sankar et al.
(1982) on this type of application in India, where artificial irrigation is used for the paddy
fields.
In these studies another advantage of the use of polymer additives in irrigation systems
was found: The dilute polymer solutions perculate slowlier through the soil than pure
water. This is due to the increased elongational viscosity of polymer solutions. The
increased elongational viscosity produces a higher resistance in the porous media flow in
the soil.
Since urea is frequently used as a fertilizer for the paddy fields, Singh et al. (1989)
developed a slow-release urea blended with the natural polymer guar-gum. Field
experiments of this blended urea were carried out in Kharagpur, India. These results
show that there may be several benefits when using drag reducing polymers in irrigation
systems in agriculture. Besides the drag reduction, which leads to an increase of the
irrigated area, the loss of water in the soil can be reduced, which is important for all those
countries in the world which suffer from a shortage of water. The polymer can be
blended with a fertilizer. Thus, this technique is promising, especially for the irrigation of
paddy fields.
APPLICATIONS 209

8.12 Possible future applications

In this chapter possible applications will be sketched, in which drag reducing additives
can be used. This includes applications which are currently being developed, i.e. drag
reducing additives which will be applied in the near future, as well as those applications
which are based on risky speculations and hypotheses. Due to the subject, this chapter of
the book will be incomplete, just as other parts of the book may be. It is the aim of this
chapter to elucidate the enormous potential of further application sites, which requires
specialized research in these fields. Perhaps these speculations will stimulate the reader's
fantasy in such a way that new solutions for well-known old problems can be found.

8.12.1 Improvement of already known applications

The first large scale application of drag reduction, the Trans-Alaska pipeline, is a good
example for showing how and why application of drag reduction started and how this
application was improved in the following years (s. Chapter 8.2). 15 years of application
- the longest permanent application - also elucidates the state of the art at the application
site. In this application as well as in other applications drag reduction was introduced in
an existing system. This means it had turned out that the capacity of an existing system
was insufficient or the energy consumption was too high. To our knowledge, so far no
permanent application exists in which the use of drag reducing agents was already
planned at the time of design. This would be the best improvement of applications
because the use of drag reducing additives later on is often hindered by details which
were not considered during the phase of design of the original installation.
The number of application sites in which surfactants are used in district heating systems,
cooling circuits, and air-conditioning will increase in the near future. The research of the
chemical industry has led to the discovery of surfactant systems which exhibit drag
reduction at surfactant systems below 100 ppm. This enables an enormous reduction of
the amount of sufactants which are needed for these closed systems and thus of the costs
for the additives. Nevertheless, it is the price of energy which has a strong influence on
the tendency to improve these installations.
The blending of polymers with a fertilizer is a promising application, especially for those
countries suffering from a shortage of water. Here, it should be mentioned that in many
of those countries the groundwater level has dropped in the last decades due to an intense
use of water for an increased agricultural production. If biopolymers are used in this type
of application, there is no danger of polluting the soil and the groundwater.
210 CHAPTER VIII

8.12.2.1 Combinations of different drag reducing systems

This book deals with drag reduction produced by the addition of three types of additives.
The description of the effect was limited to pure systems, i.e. no combinations of these
additives was considered, with one exception in Chap. 7 where the very effective
combination of fibres and polymers was discussed. Indeed, there may be several
application sites in which the combined use of two additives makes sense. In this respect
the term "combined system" should not only be restricted to these three types of additives
because drag reducing effects also occur in flows without additives.
The combined use of additives and so-called passive drag reducing systems, e.g. walls
which are coated with riblets, seems to be very promising because a synergistic effect
occurs. This synergism becomes comprehensible because the effect of riblets on
turbulence is restricted to the near-wall region where the turbulent structures and their
stability is influenced by the riblet geometry. The structure of the longitudinal vortices
becomes more pronounced and stabilized. The structures on the near-wall turbulence are
also very important for the conformation of the additives in the flow. The elongation and
the alignment of the additives is due to these structures and due to the local shear
gradient. On the other hand, these structures are mainly responsible for the degradation of
the additives. If the riblets intensify the alignment of the additives, and a reduction of
those turbulent structures which are responsible for the degradation of the additives
occurs simultaneously, a higher drag reduction will be the result.
Other combined systems are those in which the addition of soluble non-drag reducing
additives, e.g. salts, produces a change in the pH-value and/or a change in the ionic
strength of the solution. In this way the size of the dissolved molecule, its mobility and
consequently the elongation of the additive in the flow can be influenced. The chemical
stability against degradation can also be increased. In surfactant systems the temperature
range in which the solution exhibits drag reduction can be influenced by this method.
These combined systems seem to be very important for applications in biological
systems. This also holds for applications in agricultural irrigation systems as well as in
sewer systems.

8.12.2.2 Heat exchanger

The drag reduction caused by the additives reduces the turbulent momentum transport
between the flow and the wall. In general, this is accompanied by a reduction of the heat
transfer. There are applications, e.g. in crude-oil transportation (s. Chap. 8.2), in which
both effects lead to benefits. On the other hand, there are applications in which drag
APPLICATIONS 211

reduction is desired and a reduction in heat transfer should be avoided. If a separation of


these usually simultaneously occuring effects could be achieved, many new types of
application would be possible. In many unit operations of chemical engineering heat has
to be conveyed or carried off, like in central heating systems. Thus there is a need for
heat exchangers which fulfill these requirements.
Besides the method of increasing the Reynolds number and consequently the wall shear
stress above the critical in order to destroy the drag reducing micellar structure in the heat
exchanger, see Chap. 8.4, one could manipulate the flow of the additives in the heat
exchanger by other means in order to achieve a similar heat transfer as in a Newtonian
fluid. By applying external forces a support of the alignment process of the additives or a
destruction of this alignment can occur. These forces can be produced by an acceleration
of the fluid due to a rotation of the flow or by a change of the cross-section of the flow.
Furthermore, electromagnetic fields could also be used for the build-up or the destruction
of the alignment of the additives.

8.12.2.3 Drag reducing fluids in rotating flows or in flows with curved


walls

As mentioned above, a rotating field, as found in a rotating system, can influence the
structure of turbulence in flows of drag reducing fluids. In general, this holds whenever
the flow is forced to possess curved streamlines, e.g. this can be produced by curved
walls. Studies on these effects are rare in literature. It is known that Taylor-Gortler
vortices are stabilized by the addition of drag reducing additives. A theory for this
stabilization process was published by Tomita & Jotaki (1977 ).
This stabilization is not surprising because Taylor-Gortler vortices are structures of the
near-wall turbulence which are stabilized by drag reducing additives, as already known
from flows at plane smooth walls. The stabilization of Taylor-Gortler vortices is
accompanied by a reduction of the velocity fluctuations in the spanwise direction. A
possible application of this effect would be flows in centrifuges.

8.12.2.4 Jet and film flows

Besides the change in the structure of turbulence the additives also influence the surface
tension and the stability of the surface of the liquid. This is already used by several fire-
brigades. The addition of polymers reduces the friction losses in the hoses. As a
consequence, the exit velocity of the jet is increased at a constant power consumption of
the pump. Due to the stabilized surface, the reduced cross-mixing and the damping of the
small scale turbulence, the jet of a dilute polymer solution remains more compact in
212 CHAPTER VIII

comparison to the one of a Newtonian fluid at a constant Reynolds number of the jet
flow, Hoyt et al (1974) or Hoyt & Taylor (1974). This is an advantage for the fire
fighting, Thome (197 4) as well as in jet cutting, Franz (1972) and in ink jet printers, in
which additives are also used.
In the two latter applications the stability behaviour of the jet is of great importance
because the jets must have a small diameter and a spreading of the jet should be avoided.
The jet cutting is also of practical value in surgery because the jet acts as a knife and
additionally delivers the scavenging fluid.
Due to the reasons mentioned above the additives should be especially suitable for
stabilizing fast film flows. In addition, the flow-rate of the film can be increased. In
chemical engineering many processes exist in which the use of additives would be an
advantage, especially when the film flow is additionally strained, as in drying processes.

8.12.2.5 Drops and bubbles

The geometry and stability of drops and bubbles is controlled by the surface tension at
phase interfaces, especially, when additives accummulate at the interface. At interfaces
this can causes Marangoni-effects. However, the stabilization of the interfaces is much
more important. Bubbles and drops, including those in emulsions, have heat and mass
transfer with the surrounding fluid this can be due to diffusion, chemical reaction, or
phase transitions, like in the boiling or freezing of drops and these processes can be
altered by the additives.
The stabilization of the surfaces of drops is already used. Polymers are added to cerosine
as anti-misting agents in order to achieve a protection against explosion in the case of an
accident. The basic mechanism is a suppression of the atomization, similar to the
suppression of the spreading in jet flows.
The simultaneous use of the additives as drag reducers as well as stabilizing agents seems
also to be attractive: a combination of additives with bubbles and drops in a turbulent
flow. Here it can be speculated that besides a stabilization of the drops or bubbles, the
additives cause a change in the structure of the flow. This results in a changed
coalescence of the drops or bubbles. Preferably, the coalescence should be reduced,
which would result in a more homogeneous mixture. This way a new very interesting
drag reducing solution could be produced, because it is known that microbubbles are also
drag reducing, but only as long as they do not increase too much by coalescence.
APPLICATIONS 213

8.12.2.6 Laminar - turbulent transition

Depending on the type of additive, the transition laminar-turbulent will be delayed or


occur at lower Reynolds numbers. Rod-like polymers and stiff rod-like micelles shift the
transition laminar-turbulent to higher Reynolds numbers, which is in agreement with
stability calculations. The influence of flexible polymer additives on the transition laminar
-turbulent was studied by Abernathy eta!. (1980, 1984). Surprisingly, polymer-induced
velocity fluctuations were also found in laminar free-surface watertable flows, Abernathy
eta!. (1980), provided that the wall strain rate was higher than 750 s-1. Energy spectra
showed a pronounced high-frequency component in the dissipation range of a
comparable water flow, indicating that it was probably the same phenomenon as "early
turbulence". The mean velocity profile and the friction factor in such "laminar" pipe flows
containing polymers were found to be between those of a Newtonian laminar flow and
those corresponding to the maximum drag reduction asymptote, Abernathy and He
(1984).
The influence of the additives on relarninarisation is practically unknown. A drag reduced
turbulent flow could exhibit a quicker relaminarisation under some circumstances. This
would be an interesting application in cases where flows around bodies exist. On the
other hand, it has been found in pipe flows that disturbances created at the wall in
polymer solution flows persist much longer in axial direction than in Newtonian flows,
Berman & Cooper (1972). In drag reducing surfactant solutions turbulent spots exhibited
a large increase in lifetime in comparison to Newtonian fluids.

8.12.2.7 Manipulation of separation

In some applications one would like to do a manipulation of the separation. The reason
for this could be a desired reduction or an increase in drag; the latter if, for example, a
better mixing is aimed at.
In many applications it would be of interest to manipulate the separation in such a way
that the oscillation behaviour of the flow around a body could be influenced, i.e. the
Strouhal-number could be changed. The few existing publications do show an influence
of the additives on the separation effect. However, to our konwledge no application of
the effect exists nowadays. It is believed that the drag reducing slime which is produced
by some species of fish influences the separation process of the vortex at the tail of the
fish in a positive way, reducing the drag of the fish.
214 CHAPTER VIII

8.12.2.8 Entry and exit flows

The viscoelastic material properties of the fluids especially influence entry and exit flows.
In entry flows the elongational viscosity is of influence, whereas in exit flows the
relaxation of the normal stresses influences the flow. The material behaviour in these
flows is important for those types of flows in which entry and exit flows exist in series.
Here, the flow behaviour of these fluids also depends on the deformation history.
Whenever the largest relaxation time of these "memory fluids" is considerably increased,
this effect becomes important. This is known from the behaviour of polymer melts and
concentrated polymer solutions.

8.12.2.9 Exotic applications

There are many forms of application where additives can be added without having to be
extracted later on. This holds for unit operations during the food production. Here, the
natural polymers, Guar gum and Xanthan gum, according to the food laws of many
countries may be added. This is frequently used, mainly in order to increase the viscosity
of the product. Guar gum is extracted from a plant which grows in India and Pakistan,
whereas Xanthan gum is produced by bacteria in bioreactors. These additives are or
could be added during the processing of milk in large dairies, of blood in
slaughterhouses, and of fruit juices. Furthermore, these additives are also interesting for
the production of different types of food oils, instant soups, and chocolates.
Another, at the moment exotic application, would be the addition of fibres to liquid
metals. In the production of compound materials special fibres or needle crystals could
play a new important role. Here, the knowledge of fibre suspension flows is required.
Another type of application would be turbulent flows, in which adsorption properties are
of importance besides the possible drag reduction.

8.12.2.10 Application in recreation and sports

It is known that polymer gels on the skin of a swimmer or on the walls of a sailing boat
can increase efficiency twofold. Firstly, they are drag reducing and secondly they have a
positive effect on the sportsman's disposition. In this respect there are no limits to
speculation. Even, if the use of additives is forbidden in official competitions, these
additives could be added to a basin for training purposes. Besides, these additives could
be used in fun parks, where canoeing or boating courses, etc, are offered.
APPLICATIONS 215

8.12.2.11 New methods of adding

In some applications the addition of the additives is problematic. Additives can be dosed
through slits in walls or through porous walls. Here, it has to be considered that a porous
wall has a rough surface which increases the drag. If the wall were filled by a carrier
fluid, the surface would be smooth, and minute amounts of released additives would
lower the flow resistance. This method would become very attractive if the delivery
system and the flow could interact, e.g. if the porous wall were a thin membrane. The
additives would be released whenever a loc!ll low pressure field existed in the flow.
Separations with high strain fields are low pressure fields. Walls which could deliver
adsorbed additives, depending on the outer pressure field, would have a very similar
effect as porous walls. Desorbed additives would be especially elongated.
This incomplete list elucidates the variety of research areas in which the use of additives
could lead to promising applications, or, in other words, research on drag reduction is
still in progress. Many interesting and sometimes surprising effects will be found in the
future, when drag reducing additives are applied in technology.

8.13 Preparation of polymer and surfactant solutions and fibre


suspensions

In this book many material properties of polymer and surfactant solutions as well as fibre
suspensions were presented. Especially polymer solutions of extremely high molecular
weight are very sensitive to mechanical degradation (see Chap. 2.2.3). Thus it may
happen that polymer solutions are already degraded during their preparation.
Consequently, special attention has to be given to the preparation of these solutions.
Surfactant solutions have to be in a thermodynamic equilibrium in order to achieve
reproducable results in laboratory experiments. Finally, and last but not least, the flow
behaviour of fibre suspensions strongly depends on the dispersion of the fibres. Thus
this last chapter will provide some information on the preparation techniques of these
solutions or suspensions.
In order to avoid degradation in polymer solutions of high molecular weight, the
mechanical forces during the process of preparation should be as low as possible. A
simple calculation can show that the hydrodynamic forces occurring during stirring can
be orders of magnitude higher than the one in the the pipe or channel flow, in which the
behaviour of the polymer solution is studied. If the polymer solutions are stirred to
homogenize, the number of revolutions per time should be kept as low as possible. In
some cases it may become necessary to construct special tumbling machines in which a
216 CHAPTER VIII

low intensity mixing occurs. In this way the time for the preparation of a homogeneous
polymer solution increases considerably.
The polymer is normally delivered in form of a powder. Polymers which are
polyelectrolytes, e.g. polyacrylamides, exhibit the formation of gels when the powder
particles are brought in contact with water. The small solid particles of the polymer swell
and together form a gel with large local concentration gradients. If one does not avoid this
process, it becomes very difficult and time consuming to homogenize such a solution
later on. Thus special techniques have been developed in order to fasten the
homogenisation process of polyacrylamide solutions, which are frequently used in drag
reduction studies. Polyacrylamide is insoluble in higher alcohols. When the polymer
powder is suspended in isopropanol and this suspension is added to the trombe of the
stirred amount of water desired, the formation of such unwanted gels can be avoided.
However, in order to prevent the solution from degrading, the speed of the stirrer has to
be lowered as soon as, due to the increase of the viscosity, the settling velocity of the
polymer powder particles decreases considerably. In some polymer solutions, e.g.
polysaccharide or polyacrylamide solutions, algae and bacteria find good growing
conditions. (Tap water often contains seeds of algae.) This can cause a biological
degradation of the polymer solution and can be avoided by the addition ofbiocides, e.g.
formaldehyde. Polyisobutene, which is frequently studied when the behaviour of
polymers in organic solvents is studied, is a liquid of high viscosity at room temperature.
Thus in its pure form it is not available as a powder because the particles would slowly
flow and agglomerate. In technical products this can be avoided by the addition of
another powder, e.g. calciumtriphosphate. For laboratory experiments the pure polymer
can be cut into small pieces with scissors just before the preparation of the solution starts.
When surfactant solutions are prepared by dissolving the proper amount of surfactants in
the desired amount of water, it can take a long time before a thermodynamic equilibrium
in the solution is achieved. First, a broad size distribution of micelles exists. (This
solution can even exhibit better drag reducing properties because it contains some giant
micelles). In a thermodynamic equilibrium the size distribution of the micelles is
narrower. In order to accelerate the process of achieving a thermodynamic equilibrium
one can heat up the solution to a temperature at which no micelles in the solution exist any
more, i.e. at this temperature only single surfactant molecules exist in the solution. This
is a practical use of the temperature dependence of the critical micellar concentration
(CMC). If such a solution containing only single surfactant molecules is cooled down to
the desired temperature, a thermodynamic equilibrium in the solution can be achieved
sooner. However, this method is restricted to a laboratory scale. In practice, this method
is too energy consuming and thus too expensive.
APPLICATIONS 217

Due to colloidal forces, internal structures, like fibre bundles and flocks, can exist in a
suspension at rest. In order to get single, dispersed fibres in the suspension one has to
influence these colloidal forces. This can be done by adding surfactants or electrolytes.
For the preparation of the very effective drag reducing asbestos fibre suspensions
sodium palmitate or oleate as well as Aerosol OT are frequently used as surfactants,
whereas aluminiumchloride is often used as an electrolyte. In fibre suspensions, whose
fibres are very sensitive to fracture, an intensive stirring has to be avoided. Fibre bundles
can be destroyed in elongational flows, which can easily be realized in a laboratory by
contraction flows.
IX REFERENCES

Abernathy, F.H., Bertschy, J.R., Chin, R.W. & Keyes, D.E. 1980 Polymer induced
fluctuations in high strain-rate-rate laminar flows. Journal of Rheology 24,
647-665
Abernathy, F.H. & He, Z. 1984 Polymer induced velocity fluctuations in dilute drag
reducing pipe flows. In: Drag Reduction (eds. Sellin, R.H.J., Moses, R.T.),
University of Bristol, paper B8
Achia, B.U. & Thompson, D.W. 1974 Laser holographic measurement of wall-
turbulence structures in drag-reducing pipe flow. Proc. Int. Conf. Drag
reduction, Cambridge (ed. N.G.Coles), BHRA A2-23-40
Achia, B.U. & Thompson, D.W. 1977 Structure of the turbulent boundary in drag
reducing pipe flow. J. Fluid Mech. 81, 439-464
Amborn, L.& Hagstrand, U. 1977 Toms effect in district heating tube systems, Report
SVF-50, Studsvik, in Swedish
Astarita, G.& Nicodemo, L. 1970 Extensional behaviour of polymer solutions.
Chemical Engineering J. 1,57-65.
Bartram, E., Goldsmith, H.L. & Mason, S.G. 1975 Particle motions in non-Newtonian
media III. Further observations in elasticoviscous fluids. Rheol. Acta 14,
776-782
Barnes, H.A., Hutton, J.F.& Walters, K. 1989 An introduction to rheology, Elsevier,
Amsterdam
Bark, F. 1974 On the wave structure of turbulent boundary layers with application to
drag reduction. The Royal Inst. of Techn. Sweden TRITA-MEK-74-01
Bark, F.H. & Tinoco, H. 1978 Stability of plane Poiseuille flowof a dilute suspension
of slender fibres. J. Fluid Mech. 87, 321-333
Barnes, H.A. 1981 Dispersion rheology: 1980, Royal Society of Chemistry, Industrial
Division, London
Batchelor, G.K. 1953 A theory of homogeneous turbulence. Cambridge University
Press
Batchelor, G.K. 1967 An introduction to fluid dynamics. Cambridge University Press
Batchelor, G.K. 1970 The stress system in a suspension of force-free particles. J. Fluid
Mech. 41,545-570
Batchelor, G. K. 1971 The stress generated in a non-dilute suspension of elongated
particles by pure straining motion, J. Fluid Mech. 46, 813-829
Batchelor, G.K. 1977 The effect of Brownian motion on the bulk stress in a suspension
of spherical particles, J. Fluid Mech. 83, 97-117
Bechert, D.W. & Bartenwerfer, M. 1990 Turbulent drag reduction by nonplanar
surfaces-A survey on the research at TU/DLR Berlin. Structure of turbulence
and drag reduction (ed. A. Gyr) IUTAM Symp. Zurich 1989. Springer
Verl., 525-543
Beiersdorfer, H. 1994 Untersuchungen der Turbulenzstruktur bei maximaler
Widerstandsverminderung in einer Kanalstromung einer viskoelastischen
Tensidlosung. MS Thesis University of Dortmund.
Beiersdorfer, H., Bewersdorff, H.-W. & Gyr, A. 1994 Flows with surfactant at
maximum drag reduction. Proc. IUTAM Symp. On liquid particle
interactions in suspension flow. Grenoble (in print)
Bekturov, E.A. & Bakauova, Z.K. 1986 "Synthetic water-soluble polymers in
solution", Huethig & Wepf, Basel
Berman, N.S. 1978 Drag reduction by polymers. Ann. Rev. Fluid Mech. 10, 47-64
Berman, N.S. 1986 Molecular interactions in drag reduction in pipe flows. In
Encyclopedia Fluid Mech. (ed. Cherimisinoff, P.), Gulf Publ. Co, W A, 1,
1060-1081
Berman, N.S. & Cooper, E.E.l972 Stability studies in pipe flows using water and dilute
polymer solutions. AIChE J. 18, 312-320
219
220 CHAPTER IX

Berner, C. & Scrivener, 0. 1980 Drag reduction and structure of turbulence in dilute
polymer solutions. In: Viscous flow drag reduction. Progr. Astron. A
Aeronautics (ed. G.R. Hough) 72, 290-298
Bewersdorff, H.-W. 1985 Heterogeneous drag reduction in turbulent pipe flow. In: The
influence of polymer additives on velocity and temperature fields. Proc.
IUTAM Symp. Essen 1984, Springer Veri. (ed. B. Gampert), 337-348
Bewersdorff, H.W. 1990 Drag reduction in surfactant solutions. In: Structure of
turbulence and drag reduction. (ed. A.Gyr) Springer, Berlin, 293-312
Bewersdorff, H.-W. & Berman, N.S. 1987 Effect of roughness on drag reduction for
commercially smooth pipes. J. Non-Newtonian Fluid Mech. 24, 365-370
Bewersdorff, H.-W. & Berman, N.S. 1988 The influence of flow-induced non-
Newtonian fluid properties on turbulent drag reduction. Rheol. Acta 27,
130-136
Bewersdorff, H.W., Dohmann, J., Langowski, J., Lindner, P., Maack, A., Oberthi.ir,
R. & Thiel, H., 1989 SANS- and LS-studies on drag-reducing surfactant
solutions. Physica B 156 & 157 (1989) 508-511
Bewersdorff, H.W., Frings, B., Lindner, P., Oberthi.ir, R.C. 1986 The conformation of
drag reducing micelles from small-angle-neutron-scattering experiments.
Rheologica Acta 25 642-646
Bewersdorff, H.W. & Ohlendorf, D. 1988 The behaviour of drag reducing cationic
surfactant solutions. Colloid & Polymer Science 266, 941-953
Bewersdorff, H.W., Oles, V.& Martischius, F.D. 1986 Kurzfristige
KapazitiitserhOhung von Abwasserkaniilen. gwf-wasser/abwasser 127, 633-
636
Bewersdorff, H.W. & Thiel, H .. 1993 Turbulence structure of dilute polymer and
surfactant solutions in artificially roughened pipes. Appl. Sc. Res. 50, 347-
368
Bird, R.B., Armstrong, A.C., & Hassager, 0. 1987 Dynamics of polymeric liquids.
John Wiley and Sons, New York
Blackwelder, R.F. & Eckelmann, H 1979 Streamwise vortices associated with the
bursting phenomenon. J. Fluid Mech. 94, 577-594
Brandrup, J.& Immergut, E.H. 1975 Polymer handbook. Wiley Interscience, NY
Brown, D.A 1990 Attemps to achieve drag reduction via electrochemically produced
microbubbles. Proc. 2nd Int. Symp. on Performance, Enhancement for
maritime applications (ed. Nadolink, R.H.) Univ. of Rhode Island 219-226
Burger, E.D., Chorn, L.G.& Perkins, T.K. (1980): "Studies of drag reduction
conducted over broad range of pipeline conditions when flowing Prudhoe
Bay Crude Oil", Journal of Rheology 24, 603-626
Bushnell, D.M. & Hefner, J.N. 1990 Viscous drag reduction in boundary layers.
Progress in Astronautics and Aeronautics 123
Cerf, R 1951 Recherches theoriques.et experimentales sur l'effet Maxwell des solutions
de macromolecules deformable. I Theorie de l'effet Maxwell des suspensions
de spheres elastiques. J Chim Pys 48, 59-84
Chandler, R.W., Lewis, W.R. 1977 Control of sewer overflows by polymer injection.
Water Utilities Department, Dallas, Texas, U.S.A., EPA-Report 600/2-77-
189
Chang, J.C.& Denn, M.M. 1979 An experimental study of isothermal spinning of a
Newtonian and a viscoelastic liquid. Journal of Non-Newtonian Fluid
Mechanics, 5, 369-385
Chara, Z., Zakin, J.L., Severa, M. & Myska, J. 1993 Turbulence measurements of drag
reducing surfactant systems. Experiments in Fluids 16, 36-41
Chen, C.-H. P. & Blackwelder, R.F. 1978 Large-scale motion in a turbulent boundary
layer: a study ussing temperature contamination. J. Fluid Mech. 89, 1-31
Choi, K.-S. 1990 Marine application on riblets for drag reduction at high Reynolds
numbers. Proc. 2nd Int. Symp. on Performance, Enhancement for maritime
applications (ed. Nadolink, R.H.) Univ. of Rhode Island 237-244
REFERENCES 221

Coles, N.G. 1974 Drag reduction. Proc Intern. Conf in Cambridge, BHRA Fluid
Engineering
Cottrell, F.R., Merrill, E.W.&Smith, K.A 1969 Conformation of polyisobutylene in
dilute solution subjected to a hydrodynamic shear field. Journal of Polymer
Science A-2, Vol.7, 1415-1434
Cox, W.P & Merz, E.H. 1958 Correlation of dynamic and steady flow viscosities. J.
Polymer Sc. 28, 619-622
Cross, M.M. 1965 Rheology of non-Newtonian fluids: a new equation for pseudo-
plastic systems. Journal of Colloid Science 20, 412-437
Debeye, P. & Bueche, A.M. 1948 Intrinsic viscosity, diffusion, and sedimentation rate
of polymers in solution. J. Chern. Phys. 16, 573-579
Debrule, P.M.& Sabersky, R.H.1974 Heat transfer and friction coefficients in smooth
and rough tubes with dilute polymer solutions. Int. J. of Heat and Mass
Transfer 17, 529-540
de Gennes, P.O. 1976 Dynamic of entangled polymer solutions. I The Rouse model.
Macromolecules 9, 587- 593
de Gennes, P.O. 1990 General aspects of polymer chains. In: Introduction to polymer
dynamics. Cambridge University Press 1-16
Delgado, A. 1987 Untersuchung der turbulenten Stromung von Polymerlosungen in
einem zweidimensionalen Kanal mittels Laser-Doppler-Anemometrie. Ph. D.
Thesis of the GH-Essen
Den Toonder, J.M.J., Nieuwstadt, F.T.M. & Kuiken, G.D.C. 1994 The role of
elongational viscosity in the mechanism of drag reduction by polymer
additives. Appl. Sci. Res. (in print)
de Waele, A.I.C. 1923 Viscometry and plastometry. Oil Color Chern. Assoc. J. 6, 33-88
Doi, M.& Edwards, S.F. 1978 Dynamics of rod-like macromolecules in concentrated
solution, Journal of the Chemical Society, Faraday Transactions II, 74, 918-
932
Doi, M.& Edwards, S.F. 1986 The theory of polymer dynamics, Clarendon Press,
Oxford
Donohue, G.L., Tiederman, W.O. & Reischman, M.M. 1972 Flow visualization of the
near-wall region in a drag-reducing channel flow. J. Fluid Mech. 56, 559-
575
Durst, F., Haas, R. & Kaczmar, B.U. 1980 Flows of dilute HPAM-solutions in porous
media under various solvent conditions. SSB 80/E/158 Univ. of Karlsruhe
Dunlop, E.H. & Cox, L.R. 1977 Influence of molecular aggregates on drag reduction.
Phys. Fluids 20, S203-S213
Einstein, A. 1906 Eine neue Bestimmung der Molekiildimension. Ann. Physik, 19,
289-306
Elata, C. & Ippen, A.T. 1961 The dynamics of open channel flow with suspension of
neutrally buoyant particles. Tech Rep. 45 Hydrody. Lab. MIT
Elias, V.& Vocel, 1.19878 Polymer additives for sprinkler irrigation. Communications
of the Institute of Hydrodynamics 8, Prague
Elias, V.& Vocel, J. 1980a Increasing the capacity of sprinkler irrigation system network
by polymer additives. International Commision on Irrigation and Drainage,
11th Congress, 227-230
Elias, V.& Vocel, J. 1980b Long term field experiments with polymer additives for
sprinkler irrigation. Communications of the Institue of Hydrodynamics 10,
Prague
Ellis, H.D. 1970 Effect of shear treatment on drag reducing polymer solutions. Nature
226, 352- 353
Elperlin, LT., Levental, L.I.& Chesnolov, Y.N. 1971 Decreasing the hydraulic
resistance of heating networks. Thermal Engineering 18, 28-32
Flory, P.J. 1971 Priciples of polymer chemistry. Cornell University Press
Fankhanel, M. 1989 Druckverlust und Warmeiibergang in Fernwarmesystemen bei
Einsatz von mizellaren Widerstandsverminderern. Dissertation, Universitat
Dortmund
222 CHAP'IERIX

Franz, N.C. 1972 Fluid additives for improving high velocity jet cutting. Proc. 1st Int.
Symp. on jet cutting technology B.H.R.A., A7,1-11
Frenkiel, F.N., Landahl, M.T. & Lumley, J.L. 1977 Structure of turbulence and drag
reduction. Phys. Fluids 20, No 10 Part II.
Frost, W. & Moulden T.H. 1977 Handbook of turbulence 1. Plenum Press, NY.
Fuller, G.G., Cathey, C.A., Hubbard, B.& Zebrowski, B.E. 1987 Extensional
viscosity measurements for low- viscosity fluids. J. of Rheology 31, 235-
245
Fung, J.C.H., Hunt, J.C.R., Perkins, R.J., Wray, A.A. & Stretch, D. 1991 Defining
the zonal structure of turbulence using the pressure and invariants of the
deformation tensor. In "Advances in turbulence 3" (eds. A.V. Johansson &
P.H. Alfredsson) Springer Veri. Berlin 395-404
Fung, J.C.H.& Perkins, R.J. 1989 Particle trajectories in turbulent flow generated by
true-varying random Fourier modes. In "Advances in turbulence 2" (eds.
H.H. Fernholz & H.E. Fiedler) Springer Veri. Berlin 322-328
Gampert, B. 1985 The influence of polymer additives on velocity and temperature fields.
Proc. IUTAM Symp. Essen 1984, Springer Veri. Berlin, Heidelberg, NY,
Tokio
Gampert, B. & Yong, C.K. 1990 The influence of polymer additives on the coherent
structure of turbulent channel flow. Structure of turbulence and drag
reduction (ed. A. Gyr) IUTAM Symp. Zurich 1989. Springer Veri., 223-
232
Gauthier, F., Goldsmith, H.L. & Mason, S.G. 1971 The kinetics of flowing
dispersions V. Orientation distributions of cylinders in Newtonian and non-
Newtonian systems. Kolloid-Z. Z. fiir Polymere 248, 1000- 1015
Giesekus, H., Bewersdorff, H.-W., Dembeck, G., Kwade, M., Martischius, F.D. &
Scharf, R. 1981 Rheologie. Fortschritte der Verfahrenstechnik 19, 3-28
Giesekus, H., Bewersdorff, H.-W., Frings, B., Hibberd, M., Kleinecke, K., Kwade,
M., Moller, D. & Schroder, R. 1985 Rheologie. Fortschritte der
Verfahrenstechnik 23, 3-40
Giesekus, H.& Hibberd 1987 Structures of turbulence in drag reducing fluids. In
Advances in Transport Process 5 (eds. Mujumdar, A.S. & Mashelkar, R.A.)
Wiley Eastern Ltd., New Dehli, 229-284
Golda, J. 1986 Hydraulic transport of coal in pipes with drag reducing additives.
Chemical Engineering Communications 43, 53-67
Granville, P.S. 1977 Scaling-up of pipe flow frictional data for drag reducing polymer
solutions. Proc. 2nd Int. Conf. on drag reduction, BHRA (eds.Stephens,
H.S. & Clarke, J.A.) Cambridge 81,1-12
Granville, P.S. 1984 A method for predicting additive drag reduction for small diameter
pipe flow. Proc. 3rd Int. Conf. on drag reduction, BHRA (eds.Sellin,
R.H.J. & Moses, R.T.) Bristol C3,1-8
Greene, H.L., Thomas, L.C., Mostardi, R.A. & Nokes, R.F. 1974 Potential biomedical
applications of drag reducing agents. Proc. Int. Conf.on Drag reduction
Cambridge, BHRA (ed. N.G. Coles).H2-17 -27
Gust, G. 1976 Observation on turbulent-drag reduction in a dilute suspension of clay in
sea-water. J. Fluid Mech. 75, 29-47
Gyr, A. 1965 Ein Tropfenakkreszenzmodell in Atmosphare von homogen isotroper
Turbulenz. ZAMP 16,721-739
Gyr, A. 1967 The behaviour of the turbulent flow in a 2-dimensional open channel in
presence of suspended particles. Proc. 12th Int. Congr. IAHR 82, 9- 16
Gyr, A. 1968 Analogy between vortex-stretching by drag-reducing additives and vortex
stretching by fine suspensions. Nature 219, 928-929
Gyr, A. 1976 Burst cycle and drag reduction. J. appl. math. & Phys. (ZAMP) 27,717-
725
REFERENCES 223

Gyr, A. 1990 Structure of turbulence and drag reduction. Proc. IUTAM Symp.Ziirich
1989, Springer Verl. Berlin, Heidelberg, N.Y., London, Paris, Tokyo and
Hong-Kong
Gyr, A., Bewersdorff H.-W., Hoyer, K. & Tsinober, A. 1993 An investigation of
possible mechanisms of heterogeneous drag reduction in pipe and channel
flows. In "Near wall turbulent flows" Elsevier Sc. Pub. (ed. R.M.C. So,
C. G. Speziale & B.E. Launder), 679-687
Gyr, A. & MUller, A. 1975 Alteration of structures of sublayer flow in dilute polymer
solutions. Nature 253, 185-187
Gyr, A. & Schmidt, W. 1989 Stabilisation of sediment transport in pipes by drag
reducing additives. In "Drag reduction in fluid flows" Ellis Horwood Publ.
(eds R.H.J. Sellin & R.T. Moses), 223-230
Harlen, O.G. & Koch, D.L. 1993 Simple shear flow of a suspension of fibres in a dilute
polymer solution at high Deborah number. J. Fluid Mech. 252, 187-207
Heen, R. 1993 Untersuchung der Orientierung und Assoziatbildung von Tensid- und
Polymerlosungen, Dissertation, University of Dortmund
Hinch, E.J. 1976 The distortion of a flexible inextensible thread in a shearing flow. J.
Fluid Mech. 74, 317-333
Hinch, E.J. 1977 Mechanical models of dilute polymer solutions in strong flows. Phys.
of Fluids 20, S22-S30
Hinch, E.J. & Elata C. 1979 Heterogeneity of dilute polymer solutions. J. Non-
Newtonian Fluid Mech. 5, 411-425
Hinch, E.J. & Leal, L.G 1976 Constitutive equations in suspension mechanics. Part 2.
Approximate forms for a suspension of rigid particles affected by Brownian
rotations. J. Fluid Mech. 76, 187- 208
Hino, M. 1963 Turbulent flow with suspended particles. ASCE 89 Hy 4, 161- 185
Hosaka, M. 1978 Theoretical analysis of turbulent gas-solid suspension flow. J. of
Nuclear Sc. 15, 212- 218
Hough, G.R. 1980 Viscous flow drag reduction. Progress in Astronautics and
Aeronautics 72
Hoyer, K. 1994 Heterogene Widerstandsverminderung in turbulenten Rohrstromungen.
Diss ETH 10525
Hoyer, K., Bewersdorff, H.-W. & Gyr, A.1992 Studies on mechanisms of
heterogeneous drag reduction. In theoretical and applied rheology (eds. P.
Moldenaers & R. Keunings) Elsevier Amsterdam 1, 183-188
Hoyt, W.H., Taylor, J.J. & Runge, C.D. 1974 The structure of jets of water and
polymer solution in air. J. Fluid Mech. 63, 635-640
Hoyt, W.H.& Taylor, J.J. 1974 A photographic study of polymer solution: jet in air..
Proc. Int. Conf. on Drag Reduction Cambridge B.H.R.A. E 3, 1-13
Hoyt, J.W. & Sellin, R.H.J. 1991 Polymer "threads" and drag reduction. Rheol. Acta
30, 307-315
Hussain, A.K.M.F. & Reynolds, W.C. 1970 The mechanics of an organized wave in
turbulent shear flow. J. Fluid Mech. 41, 241-258
James, D.F., McLean, B.D. & Saringer, J.H. 1987 Presheared extensional flow of
dilute polymer solutions. Journal of Rheology, 31, 453-481
Jeffery, G.B. 1922 The motion of ellipsoidal particles immersed in a viscous fluid. Poe.
Roy. Soc. London A 102, 161-179
Kane, R.S., Weinbaum, S. & Pfeffer, R. 1973 Characteristics of dilute gas-solid
suspensions in drag reducing flows. Proc. 2nd Int. Conf. on pneumatic
transport of solids in pipes, BHRA Fluid Engng. Cranfield, paper C3
Kane, R.S. 1990 Drag reduction by particle addition. Viscous drag reduction in
boundary layers. (ed. Bushnell, D.M. & Hefner, J.N.) Progress in Astron.
& Aeron. 123, 433- 456
Kline, S.J. & Robinson, S.K. 1990 Turbulent boundary layer structure: progress,
status, and challenges. Structure of turbulence and drag reduction (ed. A.
Gyr) IUTAM Symp. Ziirich 1989 Springer Veri., 3-22
224 CHAPTER IX

Kuhn, W. & Kuhn, H. 1945 Bedeutung beschriinkt freier Drehbarkeit fiir die Viskositiit
und Stromungsdoppelbrechung von Fadenmolekellosungen I. Helv. Chim.
Acta 28, 1533-1579
Landahl, M.T. 1977 Dynamics of boundary layer turbulence and the mechanmism of
drag reduction. Phys. Fluids 20, S55-S63
Landahl, M.T. 1990 Hydrodynamic instability and coherent structures in turbulence.
Structure of turbulence and drag reduction (ed. A. Gyr) IUTAM Symp.
ZUrich 1989. Springer Veri., 371-397
Laufer, J. 1975 New trends in experimental turbulence research. Ann. Rev. Fluid Mech.
7, 307-326
Leal, L.G. 1975 The slow motion of slender rod-like particles in a second order fluid. J.
Fluid Mech. 69, 305-337
Leca, A.& Leca, M. 1984 Drag reduction and heat transfer measurements with
polyacraylamides on a model of a district heating system. Drag Reduction
(eds. Sellin, R.H.J., Moses, R.T.) University of Bristol D8 1-6
Lee, P.F.W. & Duffy, G.G. 1976 Relationship between velocity profiles and drag
reduction in turbulent fibre suspension flow. AIChE J. 22, 750- 753
Lee, W.K., Vaseleski, R.C. & Metzner, A.B. 1974 Turbulent drag reduction in polymer
solutions containing suspended fibres. AIChE J. 20, 128- 133
Legner, H.H. 1990 Marine applications of microbubble drag reduction. Proc. 2nd Int.
Symp. on Performance, Enhancement for maritime applications (ed.
Nadolink, R.H.) Univ. of Rhode Island 227-236
Lindner, P., Bewersdorff, H.W., Heen, R., Sittart, P., Thiel, H., Langowski, J. &
Oberthiir, R., 1990 Drag-reducing surfactant solutions in laminar and
turbulent flow investigated by small-angle neutron and light scattering.
Progress in Colloid & Polymer Science 81, 107-112
Lindgren, E.R.& Hoot, T.G.1968 Effects of dilute high molecular weight polymers on
turbulent flows of water in very rough pipe. Trans. ASME, J. Appl. Mech.
35, 417-418
Lobl, M., Thurn, H.& Hoffmann, H. 1986 Flow birefringence measurements on
viscoelastic surfactant solutions, Berichte der Bunsengesellschaft fiir
Physikalische Chemie 88, 1102- 1106
Luchik, T.S. & Tiederman, W.G. 1988 Turbulent structure in low-concentration drag-
reducing channel flows. J. Fluid Mech. 190, 241-263
Luchini, P., Manzo, F & Pozzi, A. 1991 Resistance of a grooved surface to parallel flow
and cross-flow. J. Fluid Mech. 228, 87-109
Lumley, J.L. 1964 The reduction of skin friction drag. 5th Symp. of naval
hydrodynamics, Bergen, Norway
Lumley, J.L. 1969 Drag reduction by additives. Ann. Rev. Fluid Mech. 1, 367- 384
Lumley, J.L. 1971 Some comments on the energy method. In: "Developments in
Mechanics" University of Notre Dame 6, 63-88
Lumley, J.L. 1973 Drag reduction in turbulent flow by polymer additives. J. Polymer
Sci. 7, 263-290
Lumley, J.L. 1977 Drag reduction in two phase and polymer flows. Phys. Fluids 20,
S64- S71
Lumley, J.L. & Kubo, I. 1985 Turbulent drag reduction by polymer additives: A survey
In: The influence of polymer additives on velocity and temperature fields.
Proc. IUTAM Symp. Essen 1984.(Ed. Gampert, B.) Springer Veri. Berlln,
Heidelberg, NY, Tokio 3-21
Lundgren, T.S. 1982 Strained spiral vortex model for turbulent fine structure. Phys.
Fluids 25, 2193-2203
Lycko 1993 Zum EinfluB eines kationischen Tensids auf die Turbulenzstruktur in den
Grenzschichten an einer glatten und einer rauhen Wand. University of
Dortmund
Lyons, S.L., Nikolaides, C. & Hanratty, T.J. 1988 The size of turbulent eddies close to
the wall. A. I. Ch.E. J. 34, 938-945
REFERENCES 225

Maschmeyer, R.V. & Hill, C.T. 1974 The rheology of concentrated suspensions of
fibres. Adv. Chern. Ser. 13, 95-105
Martischius, F.-D. 1982 Das rheologische Verhalten von PolymerlOsungen in Scher-
und Dehnstromungen. Rheologica Acta, 21,288-310
Martischius, F. D.& Heide, W., 1984 Drag reduction in heating systems: stabilization of
polyacrylamide solutions up to temperatures of 1500C", in: Drag reduction
(eds. Sellin, R.H.J.& Moses, R.T.), University of Bristol, D9 1-3
Matthys, E.F. 1988 Measurement of velocity for polymeric Fluids by photochromic
flow-visualization technique: the tubeless siphon. J. of Rheology 32, 773-
788
McComb, W.D. 1990 The Physics of fluid turbulence, Oxford Eng. Sc. Series 25,
Oxford: Claredon.
McComb, W.D. & Rabie, L.H. 1982 Local drag reduction due to injection of polymer
solutions into turbulent flow in_ a pipe. AIChE J. 28, 547-565
McComb, W.D. & Chan, K.T.J. 1985 Laser-Doppler anemometer measurements of
turbulent structure in drag-reducing fibre suspensions. J. Fluid Mech. 152,
455-478
McNally, W.A.1968 Heat and momentum transfer in dilute poyethylene oxide solutions.
Ph.D. Thesis, University of Rhode Island
Meissner, J. 1972 Development of a universal extensional rheometer for the uniaxial
extension of polymer melts. Trans. Society of Rheology 16, 405-420
Mih, W. & Parker, J. 1967 Velocity profile measurements and a phenomenological
description of turbulent fiber suspension pipe flow. TAPPI 50, 237- 246
Moffatt, H.K. 1969 The degree of knottedness of tangled vortex lines. J. Fluid Mech.
44, 705-719
Moffatt, H.K.l990 Fixed points of turbulent dynamical systems and suppression of
nonlinearity. In "Whither turbulence?; Turbulence at the crossroads". Lecture
notes in Physics 357, (ed. J.L. Lumley), Springer Verl. 250-257
Moffatt, H.K.1993 Spiral structures in turbulent flow. In "New approaches and
concepts in turbulence", (eds. Th. Dracos & A. Tsinober), Birkhiiuser 121-
129
Moffatt, H.K. & Tsinober, A. 1992 Helicity in laminar and turbulent flow.Annu. Rev.
Fluid Mech.24, 281-312
Moreau, J.-J. 1961 Constants d'un ilot tourbillonnaire en fluide parfait barotrope. C. R.
Acad. Sci. Paris 252, 2810-2812
Morrison, W.R.B., Bullock, K.J. & Kronauer, R.E. 1971 Experimental evidence of
waves in the sublayer. J. Fluid Mech. 47, 639-656
Motier, J.F. & Prilutski, D.J. 1984 Case histories of polymer drag reduction in crude oil
pipelines. In Sellin & Moses 1984, F2 1-14
J.F. Motier, A.M. Carrier (1989): "Recent studies on polymer drag reduction in
commercial pipelines" in: Sellin, R.H.J., Moses, R.T. (eds.) "Drag
reduction in fluid flows", Ellis Horwood, Chichester, 197-204
Moyls, A.L. & Sabersky, R.H. 1978 Heat transfer and friction coefficients for dilute
suspensions of asbestos fibers. Int. J. Heat Mass Transfer 21, 7-14
Miinstedt, H. 1975 Viscoelasticity of polysterene melts in tensile creep experiments.
Rheologica Acta 14, 1077-1088
Mysels, K.J. 1949 U.S. Patent 2, 492, 173
Nicodemo, L. & Nicolais, L 1974 Viscosity of concentrated fibre suspensions. Chern.
Engineering J. 8, 155-156
Nihoul, J.C.J. 1977 Turbulent boundary layer bearing silt in suspension. Phys. Fluids
20, S197-S202
Odell, J.A., Keller, A & Miles, M.J. 1984 Flow induced polymer degradation: Chain
halving, a new method for determining molecular weight distribution. Drag
reduction. (ed. R.H.J. Sellin & R.T. Moses) University of Bristol A.3:1-2
Offen, G.R. & Kline, S.J. (1975) A proposed model of the bursting process in turbulent
boundary layers. J. Fluid Mech. 70, 209-228
226 CHAPTER IX

Ohlendorf, D. 1985 Reibungsverminderung in turbulenten Stromungen durch Tenside.


312. Dechema Kolloquium "Viskoelastische Tensidlosungen", 17.1.1985,
Frankfurt
Oles, V. 1989 Der EinfluB widerstandsvermindernder Additive auf den FlieBvorgang in
Abwasserkanalen", Dissertation, University of Dortmund, Germany
Oliver, D.R. & Ashton, R.C. 1976 The triple jet: influence of shear on the stretching of
polymer solutions. Journal of Non-Newtonian Fluid Mechanics 1, 93-104
Olivier, D.R. & Bakhtiyarov, S.I. 1983 Drag reduction in exceptionally dilute polymer
solutions. J. Non-Newtonian Fluid Mech. 12, 113-118
Ostwald, 1925 Ueber die Geschwindigkeitsfunktion der Viskositat disperser systeme. I
Kolloid-Z. 36, 99-117
Pal, R. 1993 Pipeline flow of unstable and surfactant-stabilized emulsions. AIChE J.
39, 1754-1764
Panchev, S. 1977 Random functions and turbulence. Pergamon Press.
Perry, A.E., Schofield, W.H. & Joubert, P.N. 1969 Rough wall turbulent boundary
layers. J. Fluid Mech. 37, 383-413
Perry, A.E. & Chong, M.S. 1982 On the mechanism of wall turbulence. J. Fluid Mech.
119, 173-217
Perry, A.E. & Chong, M.S. 1987 A description of eddying motions and flow patterns
using critical-point concepts. Ann. Rev. Fluid Mech. 19, 125-155
Pfeffer, R. & Kane, R.S.1974 A review of drag reductionin dilute gas-solids suspension
flow in tubes. (ed. N.G. Coles) Proc. Int. Conf. Drag Reduction,
Cambridge BHRA, Fl-1- 21
Pollert, J.l978 The ecomomic studies for drag reduction utilization in Czechoslovakia,
CNRS Round Table, Strasbourg, not printed
Pollert, J. 1985 Today and future possibilities of industrial applications of drag
reduction. The influence of polymeric additives on velocity and temperature
fields (ed. B. Gampert), Springer, Berlin, 183-198
Pollert, J., & Kolar, V. 1977 Sludges and their transport by pipelines. Symp. of Council
of Material Economic Help. Prague, paper 17
Pollert, J., Kolar, V.& Havlik, V. 1982a Drag reduction and its engineering applications.
Acta Polytechnica, Series of Technical University in Prague, 133-142
Pollert, J.& Urbanek, M. 1982b Possibilities of decreasing friction losses in primary
heating plant distribution networks by means of polymer additives. Vodni
Hodopodastvi, Series B, 32, 275-278
Povkh, I.L., Stupin, A.B., Maksjutenko, S.N., Aslanov, P.V., Roshchin, E.A. & Tur,
A.N. 1975 Study of the turbulent flow of solutions of surface-active
materials by means of a Laser-anemometer (in Russian) Inzhenerni-
Fizicheskii Zhurnal29, 853-856
Povkh, I.L., Stupin, A.B., Maksjutenko, S.N., Aslanov, P.V. & Simonenko, A.P.
1980 Features of turbulent flows of surfactant solutions of micelle forming
surfactants (in Russian) Mekhanika turbulentnykh potokov, Tretja
vsesojusnaja Konferenzija (1977), Moscow, 44-69
Povkh, I.L., Stupin, A.V. & Aslanov, P.V. 1988 Structure of turbulence in flows with
surfactant and polymeric additives, Fluid-Mechanics- Soviet Research 17,
65-79
Radin, 1., Zakin, J.L. & Patterson, G.K. 1975 Drag reduction in solid-fluid systems.
AIChE J. 21, 358- 371
Rehage, H., Wunderlich, 1.& Hoffmann, H. 1986 Progress in Colloid & Polymer
Science 72, 51
Reischman, M.M. & Tiederman, W.G. 1975 Laser-Doppler anemometer measurements
in drag reducing channel flows. J. Fluid Mech. 70, 369-392
Riediger, S. 1988 The influence of drag reducing additives on the coherent structures in
a face shear layer. 2nd Europ. Turb. Conf.,Berlin, see also Proc. 6th Symp.
Turb. Shear Flows, Toulouse 1987, 14.5.1-14.5.2
Robertson, A.A. & Mason, S.G. 1957 The flow characteristics of dilute fiber
suspensions. T APPI 40, 326- 334
REFERENCES 227

Robinson, S.K. 1990 A review of vortex structures and associated coherent motions in
turbulent boundary layers. Structure of turbulence and drag reduction (ed. A.
Gyr) IUTAM Symp. Ziirich 1989. Springer Verl., 23-50
Robinson, S.K. 1991 The kinematics of turbulent boundary layer structure. NASA
Tech. Memorandum 103859, 479 pages
Rose, G.D., Foster, K.L., Slocum, V.L.& Lenhart, J.G. 1984 Drag reduction and heat
transfer characteristics of viscoelastic surfactant formulations. In: Drag
Reduction (eds. Sellin R.H.J., Moses, R.T.), University of Bristol D61-7
Rouse, P.E. 1953 A theory of linear viscoelastic properties of dilute solutions of coiling
polymers. J. Chern. Phys. 21, 1272-1280
Sankar, G., Singh, R.P.& Singh, J. 1982 Application of drag reducing polymers in
reducing energy requirements of sprinkler irrigation systems II. Journal
Agricultural Engineering 19,9-14
Savins, J.G. 1967 A stress-controlled drag-reduction phenomenon. Rheolgica Acta 6,
323- 330
Schmid, A. 1985 Wandnahe turbulente Bewegungsablaufe und ihre Bedeutung fiir die
Riffelbildung. Ph. D. Thesis ETHZ Nr. 7697
Schmitt, W. 1989 Riffelbildung und Sedimenttransport in Rohrstromungen in
Abhangigkeit der Dichte des Sohlenmaterials fiir Reinwasser und
Polymerlosungen. Ph. D. Thesis Nr. 8895
Schiimmer, P. & Thielen, W. 1980 Structure of turbulence in viscoelastic fluids. Chern.
Eng. Commun. 4, 593-606
Sellin, R.H.J. 1977 Increasing sewer capacity by polymer dosing. Proc. Inststn. Civ.
Engrs. 63, part 2, 49-67
Sellin, R.H.J. 1978 Drag reduction in sewers: First results from a permanent
installation. Journal of Hydraulic Research 16, 357-371
Sellin, R.H.J. 1980 Polymer drag reduction in large pipes and sewers. Results of recent
field trials. Journal of Rheology 24,667-684
Sellin, R.H.J.,Hoyt, J.W. & Scrivener, 0. 1982 The effect of drag reducing additives
on fluid flows and their industrial application. Part 1: Basic aspects, J. Hydr.
Res. 20, 29-68. Part 2: Present applications and future proposals. J. Hydr.
Res. 20, 235-292
Sellin, R.H.J. & Moses, R.T. 1984 Drag reduction. Proc.lntern. Conf. in Bristol,
University of Bristol
Sellin, R.H.J. & Moses, R.T. 1989 Drag reduction in fluid flows. Proc.lntern. Conf. in
Davos, IAHR, Ellis Horwood, Chichester
Shaqfeh, E.S.G. & Koch, D.L. 1990 Orientational dispersion of fibers in extensional
flows. Phys. Fluids A 2, 1077- 1093
Singh, R.P., Singh, J. & Holey, S.A.1979 Application of drag reducing polymers in
reducing energy requirements of sprinkler irrigation systems I, Journal
Agricultural Engineering 16, 53-59
Singh, R.P., Singh, J., Kumar, D., Kumar, A. & Deshmukh, S.R. 1989 Novel
applications of drag reducing polymers in agriculture. In: Drag reduction in
fluid flows (eds. Sellin, R.H.J., Moses, R.T.) Ellis Horwood, Chichester,
240-246)
Sittart, P. & Bewersdorff, H.-W. 1989 Upscaling in heterogeneous drag reduction
systems. In: Proc 3rd European Rheology Conf., (ed. D.R. Oliver) Elsevier,
London, 450-452
Smith, C.R., Walker, J.D.A., Haidari, A.H. & Taylor, B.K. 1990 Hairpin vortices in
turbulent boundary layers: The implications for reducing surface drag. In:
Structure of turbulence and drag reduction (ed. A. Gyr) IUTAM Symp.
Ziirich 1989. Springer Verl., 51-58
Smith, R.E. & Tiederman, W.G. 1991 The mechanism of polymer thread drag
reduction. Rheol. Acta 30, 103-113
Spangler, J.G.(l969) Studies of viscous drag reduction with polymers including
turbulence measurements and roughness effects. In: "Viscous Drag
Reduction" (ed. C.S. Wells), 131-157
228 CHAPTER IX

Steiff, A., Althaus, W., Weber, M.& Weispach, P.M. 1989 Applications of drag
reducing additives in district heating systems - present state of investigations.
Drag reduction in fluid flows (eds. Sellin, R.H.J., Moses, R.T.), Ellis
Horwood, Chichester, 247-254
Stephens, H.S. & Clarke J.A. 1977 Drag reduction. Proc.lntern. Conference in
Cambridge, BHRA Fluid Engineering
Tanner, R.I. 1985 Engineering Rheology. Oxford University Press, Oxford
Taylor, G.I. 1960 Conditions at the surface of a hot body exposed to the wind.
Cambridge Univ. Press, Scientific papers, II, 27-32
Tennekes, H. 1966 Wall region in turbulent shear flow of non-Newtonian fluids ..Phys.
Fluids 9, 872-878
Tennekes, H. & Lumley, J.L. 1972, A first course in turbulence. MIT Press.
Thorne, P.F. 1974 Drag reduction in fire-fighting. Proc. Int. Conf on Drag reduction
Cambridge. B.H.R.A. HI, 1-16
Tiederman, W.G. 1974 A contribution on the effect on drag reduction upon flow in the
near-wall region. Proc. Int. Conf. Drag reduction, Cambridge (ed.
N.G.Coles), BHRA A79-81
Tirtantmadja, V. & Sridhar, T. 1993 A filament stretching device for measurements of
extensional viscosity. J. of Rheology 35, 1081-1102
Tomita, Y. & Jotaki, T. 1977 Effects of elongational viscosity of polymer solution on
Taylor- Gortler vortices. Phys. Fluids 20, S75-S77
Toms, B.A. 1948 Some observations on the flow of linear polymer solutions through
strait tubes at large Reynolds numbers. North Holland, Amsterdam, Proc.
1st Intern. Congr. on Rheology 2, 135-141
Trouton, F.T. 1906 On the coefficient of viscous traction and its relation to that of
viscosity. Proc. Royal Soc. A77, 426-440
Tsinober, A. 1989 On one property of Lamb vector in isotropic turbulent flow. Phys.
Fluids A 2, 484-486
Tsinober, A. 1990 Turbulent drag reduction versus structure of turbulence. Structure of
turbulence and drag reduction (ed. A. Gyr) IUTAM Symp. ZUrich 1989.
Springer Veri., 313-340
Tsinober, A. 1990 MHD flow drag reduction. Viscous drag reduction in boundary
layers. (ed. Bushnell, D.M. & Hefner, J.N.) Progress in Astron. & Aeron.
123, 327-349
Usui, H. 1990 Drag reduction caused by the injection of a polymer solution into a pipe
flow. Structure of turbulence and drag reduction (ed. A. Gyr) IUTAM
Symp. ZUrich 1989. Springer Verl., 257-274
Usui, H., Sakuma, Y. & Saeky, T. 1993 Reynolds stress defect in polymer drag
reducing flow. 9th Symposium "Turbulent Shear Flow" Kyoto, 20-3, 1-6
Virk, P.S, 1971 Drag reduction in rough pipes. J. Fluid Mech. 45, 225-246
Virk, P.S. 1975 Drag reduction fundamentals. AIChE 21, 625-656
Virk, P.S. & Wagger, D.L. 1990 Aspects of mechanisms in type B drag reduction.
Structure of turbulence and drag reduction (ed. A. Gyr) IUTAM Symp.
ZUrich 1989. Springer Verl., 201-213
Vissmann, K.& Bewersdorff, H.-W. 1990 The influence of pre-shearing on the
elongational behaviour of dilute polymer and surfactant solutions. Journal of
Non-Newtonian Fluid Mech., 34, 289-317
Vrahopoulou, E.P. & McHugh, A.J. 1987 Shear-thickening and structure formation in
polymer solutions. Journal of Non-Newtonian Fluid Mech., 25, 157-175
Wade, R.H. 1972 Symposium on polymer reduction, New York, U.S.A., paper 13 c
Western Company 1969 Polymers for sewer control, Report no. WP-20-22,U.S.
Federal Water Pollution Control Administration, Richardson, Texas, U.S.A.
Wilmarth, W.W., Wei, T. & Lee, C.O. 1987 Laser anemometer measurements of
Reynolds stresses in a turbulent channel flow with drag reducing polymer
additives. Phys. Fluids, 30, 933-935
REFERENCES 229

Wolff, C.l982 Non-Newtonian behaviour of associations of macromolecules in dilute


solutions. Advances in Colloid and Interface Science, 17, 263-274
Wray, A.A. & Hunt, J.C.R. ( 1990) Algorithms for classification of turbulent structures.
In "Topological fluid mechanics", (eds H.K. Moffatt & A. Tsinober) 95-104
+ one colour plate
Wunderlich, A.M. & James, D.P. 1987 Extensional flow resistance of dilute
polyacrylamide and surfactant solutions. Rheologica Acta, 26, 522-531
Yong, C.K. 1990 Zur Wirkung von Polymer-Additiven auf die kohiirente Struktur
turbulenter Kanalstromungen. Ph. D. Thesis at the University of Essen.
Zakin, J.L., Ni, C.C., Hansen, R.J. & Reischman, M.M. 1977 Laser Doppler
velocimetry studies of early turbulence. Phys. Fluids 20, S85-S88
Zimm, B.H. 1956 Dynamics of polymer molecules in dilute solutions: vciscoelasticity,
flow birefringence and dielectric loss. J. Chern. Phys. 24, 269-278
X INDEX
A Couette flow, 3, 40
active, 129 viscometer, 41
additive-containing paints, 202 Cox-Merz-rule, 51
agglomerates, 128 critical points, 124
algae, 208 Cross model, 58
alignment, 145, 147 cross-section, 8
anionic, 27
anisotropic flow structure, 148 D
anti-thixotropic, 35 d-type roughness, 131
apparent roughness, 132 Deborah number, 36, 184, 189
asbestos fibres, 179, 192 deformation history, 34
aspect ratio, 178 rate, 60
association number, 28 degradation, 24, 194
asymptotic friction factor, 11 deviatoric stress tensor, 70
autocorrelation, 84 die swell effect, 37
Avogadro's number, 152 dilute, 18, 57
discharge, 8
B discs, 28
Beltrami flow, 75, 148 dissipation, 154
biological degradation, 216 mechanism, 144
blades, 205 DNA molecules, 207
blood pressure, 207 dosing, 196
Boltzmann constant, 21, 141 drag reduction, 8
Brownian motion, 22, 125 dumb-bell model, 19
buffer layer, 13 duration of the event, 121
zone, 109 dynamic shear viscosity, 34
bulk velocity, 8 dynamic viscosity, 37
Burgers model, 55 E
burst cycle, 145 eddies, 124
period,93 eddy-fibre length scale, 186
effective viscosity, 113, 119
c effectively smooth, 132
capillary viscometers, 44 ejections, 93, 120
cascade process, 154 elastic constant., 19
cationic, 27 dumbbell model, 189
cavitation, 205 ellipsoids, 23
change in energy, 72 limit, 142
characteristic flow zones, 122 rigid body, 34
circulation, 7 4 stresses, 112
clay, 206 sublayer, 153
coalescence, 212 zone, 13, 101, 103
coherent structures, 76, 92, 120, 122 elongated additives, 147
coil-stretch transition, 189 elongational flows, 39
concentrated, 57 rate, 39
cone-and plate-viscometer, 43 viscosity, 39, 145
conservation laws, 70 energy budget, 78
of momentum, 70 dissipation, 81, 86
consistency, 58 equation, 78
constitutive equation, 16, 54 transport equation, 78
continuity equation, 70 enstrophy,81, 148
convergence fields, 124 generation, 82
correlation coefficient, 83 entanglements, 126
231
232 INDEX

environment, 191 integral scales, 84


equation of continuity, 77 intensity of an event, 121
equilibrium range, 90 interparticle separation, 152
extended length, 152 intrinsic slope increment, 103
extensional, 39 viscosity, 48
irrigation, 408
F isolated, 126
Fanning's friction factor, 8 isotropic, 114
Fano-flow, 46 assumption, 98
fibre suspensions, 215 isotropy, 114, 148
fibre-like flow regime, 183
fibre/polymer mixture, 185 J
fibrous suspensions, 178 .Jeffrey fluid, 56
finite extensibility, 22 orbits, 184
fixed costs, 191 jet-like, 124
flatness, 117 joint probabilities, 88
factor, 87 probability density function, 113
Flory radius., 21
flow rate, 8 K
food production, 214 Kelvin model, 55
four quadrants, 115 kinematic approach, 122
fourth moment, 87 viscosity, 10, 72
fractional slip, 132 Kolmogorov scale, 86, 149
friction factor, 8, 101 time scale, 87
fully turbulent region, 180 velocity, 90
kurtosis, 87, 117
G
global friction factor, 19 L
globular, 28 Lamb-vector, 149
lateral, 83
H length scale, 84
hairpin vortices, 146 micro scale, 86
head-group, 27 law of the wall, 101
heat transfer, 190 length scale ratio, 138
heating systems, 200 local isotropy, 90
helicity, 75 micro flow fields, 122
density, 75, 150 logarithmic layer, 13
heterogeneous drag reduction, 126, 127 profile, 101
solutions, 127 longitudinal, 83
higher moments, 117 micro scale, 86
Hooke's Law, 34 length scale, 84
Hookean solid, 34 vortices, 93
hydrodynamic frequency, 140 loss modulus, 37
radius, 18
hydrofoil, 203 M
hydrophilic, 27 macromolecule contour length, 152
hydrophobic, 27 maintenance, 191
hydrotransport, 204 Mark-Houwink exponent, 48
maximum drag reduction asymptote, 101
I MDR asymptote, 7
incompressible liquids, 70 Maxwell model, 55
inertial subrange, 91 mean rate of strain, 79
inextensible flexible thread, 23 velocity, 153
micelles, 28, 157
INDEX 233

micro bubbles, 189, 203 Prandtl-Karman coordinates, 101


micro-length scale, 86 law, 101
micro-time scale, 85 preshearing effect, 134
mixed flow region, 180 pressure drop, 9
mixing length concept, 120 fluctuations, 109
length constant, 103 loss, 101
molecular weight, 17 propellers, 205
monodisperse, 26, 126 pumping station, 195
monomer size, 21
monomeric, 17 Q
motion of vorticity, 73 0-solvent, 48

N R
Navier-Stokes equation, 70, 71, 77 Rabinowitsch correction, 44
Newton's Law, 34 radius of gyration, 18
Newtonian fluid, 34 rate of dissipation, 72
Newtonian fluids, 4 reduced viscosity, 47
plug, 103 reflection, 121
no-slip condition, 74 relative viscosity, 47
noise, 205 relaxation time, 20, 36, 125, 127, 138
nonionic, 27 retro-onset, 8, 104, 106, 151
normal stress difference, 39 Reynolds equation, 77
stresses, 38 stress deficit, 112
number concentration, 126 stress tensor, 77
density, 18 rheological behaviour, 16
rheology, 16
0 rheopectic, 35
onset, 101, 153 riblets, 203
constant, 102 rigidity modulus, 34
phenomenon, 6 rod-like, 28
point, 151 rotation tensor, 60
value, 101 roughness elements, 132
wall shear stress, 102
Ostwald-de-Waele fluid, 58 s
over-structures, 30 sand roughnesses, 131
overall production modul, 146 suspensions, 175
overlap concentration, 57 second critical concentration, 30
moment, 87
p sediment transport, 206
paperpulp, 175 semi-dilute, 57
parallel-plate viscometer, 43 separation process, 192
percent drag reduction, 9 sewer channel, 206
phase volume, 65 shear induced state, 158
plug flow, 180 layers, 124
Poisson solution, 78 rate, 34
polydisperse, 26 stress, 34, 38
polyelectrolytes, 27 thinning, 35, 61, 105, 159
polymer aggregates, 127 thickening, 35
solutions, 215 ship hull, 202
power spectral density, 90 skewness, 87, 117
spectrum, 89, 117 skin friction, 8
power-law index, 58 slime, 207
power-law-fluid, 58 slope increment, 13, 101, 102, 151
small-angle-neutron-scattering, 157
234 INDEX

solvent density, 152 v


viscosity, 20 valves, 205
space correlation, 83 vapor pressure, 205
space-time correlation, 84, 88 variable costs, 191
spectrum of dissipation, 91 variance, 87
spongy ball, 19 velocity fluctuation, 97
spotty interaction, 95 Virk's MDR asymptote, 11
storage modulus, 37, 142 ultimate profile, 13, 101, 128
strain, 33 viscoelastic fluid, 4, 34
rate, 34, 60, 79 solid, 34
streaky pattern, 122 viscometer, 40
streaming, 124 viscosity, 34
stretched, 145 viscous dissipation, 79
stretching limit, 142 forces, 73
structure of turbulence, 16 sublayer, 13, 101, 121
surfactant solutions, 215 volume fraction, 47
surfactant-stabilized emulsion, 189 von Karman constant, 181
sweeps, 93, 120 vortex stretching, 74
vorticity, 73
T equation, 75
Taylor micro scale, 86, 149 vortons, 124
Taylor's hypothesis., 84
series, 85 w
tension-thickening, 40 wall coordinates, 10
tension-thinning, 40 shear stress, 8
third moment, 87 wave number spectral density, 88
thixotropic, 35 range, 186
three-dimensional Fourier transform, 88 Weisenberg effect, 4
wave number vector, 88
threshold concentration, 143 X
time correlation, 84 Xanthan gum solutions, 106
probability density, 115
scale ratio, 138 z
topological aspects, 122 zero-shear viscosity, 48, 57
towing tank, 203 Zimm expression, 21
transition concentration, 30 relaxation rate, 140
Trouton ratio, 40, 71
turbulence intensity, 109
production, 79, 144, 154
two-dimensional, 148
two-loci method, 106
type A drag reduction, 150
type B drag reduction, 150

u
universal velocity profile, 13
unsteadiness, 97
up-scaling problem, 106
Mechanics
FLUID MECHANICS AND ITS APPLICATIONS
Series Editor: R. Moreau
Aims and Scope of the Series
The purpose of this series is to focus on subjects in which fluid mechanics plays a fundamental
role. As well as the more traditional applications of aeronautics, hydraulics, heat and mass transfer
etc., books will be published dealing with topics which are currently in a state of rapid develop-
ment, such as turbulence, suspensions and multiphase fluids, super and hypersonic flows and
numerical modelling techniques. It is a widely held view that it is the interdisciplinary subjects that
will receive intense scientific attention, bringing them to the forefront of technological advance-
ment. Fluids have the ability to transport matter and its properties as well as transmit force,
therefore fluid mechanics is a subject that is particularly open to cross fertilisation with other
sciences and disciplines of engineering. The subject of fluid mechanics will be highly relevant in
domains such as chemical, metallurgical, biological and ecological engineering. This series is
particularly open to such new multidisciplinary domains.

1. M. Lesieur: Turbulence in Fluids. 2nd rev. ed., 1990 ISBN 0-7923-0645-7


2. 0. Metais and M. Lesieur (eds.): Turbulence and Coherent Structures. 1991
ISBN 0-7923-0646-5
3. R. Moreau: Magnetohydrodynamics. 1990 ISBN 0-7923-0937-5
4. E. Coustols (ed.): Turbulence Control by Passive Means. 1990 ISBN 0-7923-1020-9
5. A.A. Borissov (ed.): Dynamic Structure of Detonation in Gaseous and Dispersed Media. 1991
ISBN 0-7923-1340-2
6. K.-S. Choi (ed.): Recent Developments in Turbulence Management. 1991
ISBN 0-7923-1477-8
7. E.P. Evans and B. Coulbeck (eds.): Pipeline Systems. 1992 ISBN 0-7923-1668-1
8. B. Nau (ed.): Fluid Sealing. 1992 ISBN 0-7923-1669-X
9. T.K.S. Murthy (ed.): Computational Methods in Hypersonic Aerodynamics. 1992
ISBN 0-7923-1673-8
10. R. King (ed.): Fluid Mechanics of Mixing. Modelling, Operations and Experimental Tech-
niques. 1992 ISBN 0-7923-1720-3
11. Z. Han and X. Yin: Shock Dynamics. 1993 ISBN 0-7923-1746-7
12. L. Svarovsky and M.T. Thew (eds.): Hydroclones. Analysis and Applications. 1992
ISBN 0-7923-1876-5
13. A. Lichtarowicz (ed.): Jet Cutting Technology. 1992 ISBN 0-7923-1979-6
14. F.T.M. Nieuwstadt (ed.): Flow Visualization and Image Analysis. 1993 ISBN 0-7923-1994-X
15. A.J. Saul (ed.): Floods and Flood Management. 1992 ISBN 0-7923-2078-6
16. D.E. Ashpis, T.B. Gatski and R. Hirsh (eds.): Instabilities and Turbulence in Engineering
Flows. 1993 ISBN 0-7923-2161-8
17. R.S. Azad: The Atmospheric Boundary Layer for Engineers. 1993 ISBN 0-7923-2187-1
18. F.T.M. Nieuwstadt (ed.): Advances in Turbulence IV. 1993 ISBN 0-7923-2282-7
19. K.K. Prasad (ed.): Further Developments in Turbulence Management. 1993
ISBN 0-7923-2291-6
20. Y.A. Tatarchenko: Shaped Crystal Growth. 1993 ISBN 0-7923-2419-6

Kluwer Academic Publishers - Dordrecht I Boston I London


Mechanics
FLUID MECHANICS AND ITS APPLICATIONS
Series Editor: R. Moreau
21. J.P. Bonnet and M.N. Glauser (eds.): Eddy Structure Identification in Free Turbulent Shear
Flows. 1993 ISBN 0-7923-2449-8
22. R.S. Srivastava: Interaction of Shock Waves. 1994 ISBN 0-7923-2920-1
23. J.R. Blake, J.M. Boulton-Stone and N.H. Thomas (eds.): Bubble Dynamics and Interface
Phenomena. 1994 ISBN 0-7923-3008-0
24. R. Benzi (ed.): Advances in Turbulence V. 1995 ISBN 0-7923-3032-3
25. B.I. Rabinovich, V.G. Lebedev and A.I. Mytarev: Vortex Processes and Solid Body Dynamics.
The Dynamic Problems of Spacecrafts and Magnetic Levitation Systems. 1994
ISBN 0-7923-3092-7
26. P.R. Voke, L. Kleiser and J.-P. Chollet (eds.): Direct and Large-Eddy Simulation I. Selected
papers from the First ERCOFfAC Workshop on Direct and Large-Eddy Simulation. 1994
ISBN 0-7923-3106-0
27. J.A. Sparenberg: Hydrodynamic Propulsion and its Optimization. Analytic Theory. 1995
ISBN 0-7923-3201-6
28. J.P. Dijksman and G.D.C. Kuiken (eds.): IUTAM Symposium on Numerical Simulation of
Non-Isothermal Flow of Viscoelastic Liquids. Proceedings of aR IUTAM Symposium held in
Kerkrade, The Netherlands. 1995 ISBN 0-7923-3262-8
29. B.M. Boubnov and G.S. Golitsyn: Convection in Rotating Fluids. 1995 ISBN 0-7923-3371-3
30. S.I. Green (ed.): Fluid Vortices. 1995 ISBN 0-7923-3376-4
31. S. Morioka and L. van Wijngaarden (eds.): IUTAM Symposium on Waves in Liquid/Gas and
Liquid/Vapour Two-Phase Systems. 1995 ISBN 0-7923-3424-8
32. A. Gyr and H.-W. Bewersdorff: Drag Reduction ofTurbulent Flows by Additives. 1995
ISBN 0-7923-3485-X

Kluwer Academic Publishers - Dordrecht I Boston I London

You might also like