Sie sind auf Seite 1von 130

Rheology and structure

of COMPLEX FLUIDS

Jan Vermant
Department of Chemical Engineering
K.U. Leuven

Lecturer : Jan Vermant

The material in this lecture is covered by some textbooks, some of them however
being too limited, other ones being probably too detailed. For those among you who
want to know more about rheology we advice the following books:
Introduction to Rheology (Rheology Series) by H.A. Barnes, J.F. Hutton, K.
Walters, Elsevier Science Ltd (1999-2nd edition)
Rheology : Principles, Measurements, and Applications (Advances in
Interfacial Engineering) by Christopher W. Macosko, Wiley-VCH (1994)
The Structure and Rheology of Complex Fluids (Topics in Chemical
Engineering) by Ronald G. Larson , Oxford Univ Press (1999)

For those of you who want to see what is available on the WWW, good starting
places are:
The rheology index, a list of rheology related web sites
http://www.rheology.org/sorindex/default.asp
The European Society of Rheology : http://www.rheology-esr.org/

1
What is Rheology??

= flow

Rheology = The Science of Deformation and Flow


1920 : Eugene Bingham (Lafayette College)

Why do we need it?


- measure fluid properties
- understand structure-property relations
- model behaviour
- to predict flow behaviour
of complex liquids
under processing conditions

Rheology is defined as the science of deformation and flow. In principle, this


definition includes everything that deals with flow, such as fluid dynamics, hydraulics,
aeronautics and even solid state mechanics. However, in rheology we tend to focus
on materials that have a deformation behaviour in between that of liquids and solids,
i.e. with a funny material behaviors. The science of rheology started in the 1920s
when polymers started to be produced, leading to novel polymeric substances and
new colloidal fluids (e.g. paints).

Hence the Newtonian fluid and the Hookean elastic solids are outside the scope of
rheology and material behaviour intermediate to these classical extremes will be
studied here. The term viscoelastic is used to describe this behavior. Some fluids
are however essentially inelastic, but have a viscosity which changes with the
deformation state, they are called Non-Newtonian fluids.

Rheology is an applied science, and its aim is twofold: Firstly, rheologists try to
understand the relation between structure and flow properties.This important for the
intelligent design and/or formulation of materials for certain applications. Secondly, by
studying the material behavior using simple deformations, fundamental relations will
be derived between deformation and force. Such equations are called constitutive
equations. These equations can then be used to predict the material behavior in
complex deformation histories as they take place in typical process operations: e.g.
extrusion, polymer film blowing, spraying, pumping etc.

2
1. Viscoelasticity
I - Silly Putty
G(t)

t< t>

Polydimethylsiloxane + Borax
Dr. James Wright, a GE engineer, came upon the material by mixing silicone oil with boric acid.

Peter Hodgson borrowed $147, bought the production rights from GE, and began producing the goo.
He renamed it Silly Putty, and packaged it in plastic eggs because Easter was on the way.
Peter Hodgson's product left him an estate of $140 million at his death in 1976.

Probably the best-known example of a material with a funny material is the so-called
Bouncing or silly putty. When this material is rolled into a small ball and dropped on
a surface, it will bounce. When it left to sit on a flat surface it will flow or like any liquid
will take on the shape of the container that holds it. Depending on the time scale on
which the deformations take place, the material will behave like either a solid (short
times) or a liquid (long times). But most of the the time the material will show an
intermediate behavior, displaying characteristic of a solid and a liquid at the same
time. We will see later that we can describe this by a time-dependent modulus G(t).
We will also learn how we can quantify to which extent a material is fluid or solid-like,
by means of linear viscoelastic measurements.

Like the modulus, the viscosity is also not always a constant. Three examples are
given above. Mayonnaise is an example of a shear-thinning fluid. When it sits on a
knife the only forces exerted on it are gravitational (low stresses) and the viscosity is
high. When the mayonnaise is applied to a sandwich, it is being sheared at high
stresses and the viscosity is low. The viscosity will be dependent on the rate at which
the material is being deformed, and decreasing as the shear rate is increased. This is
a behavior also typically encountered in polymers. Ketchup is an example of a
material which will not flow below a certain stress level, called a yield stress. It also
has a viscosity which, in addition is strongly dependent on the deformation history and
changes its value in time, this phenomenon is called thixotropy and is sometimes
encountered in suspensions. Concentrated dispersions can exhibit even different
features : a viscosity which increases with the deformation rate. This behaviour called
shear-thickening.

3
2. Non-Newtonian behaviour
II - Mayonnaise

- Ketchup "(!! , t )
- Starch

- Liquid Body armor

http://www.ccm.udel.edu/STF/index.html

Probably the best-known example of a material with a funny material is the so-called
Bouncing or silly putty. When this material is rolled into a small ball and dropped on
a surface, it will bounce. When it left to sit on a flat surface it will flow or like any liquid
will take on the shape of the container that holds it. Depending on the time scale on
which the deformations take place, the material will behave like either a solid (short
times) or a liquid (long times). But most of the the time the material will show an
intermediate behavior, displaying characteristic of a solid and a liquid at the same
time. We will see later that we can describe this by a time-dependent modulus G(t).
We will also learn how we can quantify to which extent a material is fluid or solid-like,
by means of linear viscoelastic measurements.

Like the modulus, the viscosity is also not always a constant. Three examples are
given above. Mayonnaise is an example of a shear-thinning fluid. When it sits on a
knife the only forces exerted on it are gravitational (low stresses) and the viscosity is
high. When the mayonnaise is applied to a sandwich, it is being sheared at high
stresses and the viscosity is low. The viscosity will be dependent on the rate at which
the material is being deformed, and decreasing as the shear rate is increased. This is
a behavior also typically encountered in polymers. Ketchup is an example of a
material which will not flow below a certain stress level, called a yield stress. It also
has a viscosity which, in addition is strongly dependent on the deformation history and
changes its value in time, this phenomenon is called thixotropy and is sometimes
encountered in suspensions. Concentrated dispersions can exhibit even different
features : a viscosity which increases with the deformation rate. This behaviour called
shear-thickening.

4
III. Weissenberg effect

N1

A spectacular example of elasticity in an fluid is the so-called Weissenberg or rod-


climbing effect. When a rod is set to rotate in a Newtonian fluid, the inertial forces
push the material to the outside of the container. In a visco-elastic fluid that is
subjected to deformation stresses will develop along the normal axes of the flow field.
In the case of a rotating rod, the stream lines are circular. If the normal stress in the
circular direction of the flow is greater in magnitude than the two mutually
perpendicular components, a tension in the flow direction results. These circular
normal stresses that arise will strangle the material around the rod, pushing it up. The
above photograph was prepared by dissolving a high molecular weight Polyisobutene
(Oppanol B200) in a low molecular weight sample of the same chemical composition
(From Boger and Walters, Rheological Phenomena in focus,Elsevier ,1993).

Die-swell of a fluid is another phenomenon which is typical for viscoelastic fluids. The
above photos compare the behavior of a Newtonian material with a polymeric fluid
exciting from a die. The ratio of the diameter of the jet to the capillary diameter is
113% for the Newtonian case (low Reynolds number) whereas it can be as high as a
few 100% for the case of polymeric fluids. This is a very important phenomenon when
extruding products. The die-swell will need to be accounted for (think about how you
would make a square duct by extrusion). Again the die swell arises mainly because of
differences in the normal stresses along and perpendicular to the stream lines in the
tube. They relax as the fluid exists from the tube causing a contraction in the
longitudinal direction. In the above pictures the Reynolds number is 0.001. The fluid
on the top is a Newtonian oil, the second fluid is an elastic fluid prepared in the same
manner as discussed previously (From Boger and Walters, Rheological Phenomena
in focus (Elsevier 1993).

5
IV. Die Swell

Weissenberg effect

McKinley et al (MIT)

N1

A spectacular example of elasticity in an fluid is the so-called Weissenberg or rod-


climbing effect. When a rod is set to rotate in a Newtonian fluid, the inertial forces
push the material to the outside of the container. In a visco-elastic fluid that is
subjected to deformation stresses will develop along the normal axes of the flow field.
In the case of a rotating rod, the stream lines are circular. If the normal stress in the
circular direction of the flow is greater in magnitude than the two mutually
perpendicular components, a tension in the flow direction results. These circular
normal stresses that arise will strangle the material around the rod, pushing it up. The
above photograph was prepared by dissolving a high molecular weight Polyisobutene
(Oppanol B200) in a low molecular weight sample of the same chemical composition
(From Boger and Walters, Rheological Phenomena in focus,Elsevier ,1993).

Die-swell of a fluid is another phenomenon which is typical for viscoelastic fluids. The
above photos compare the behavior of a Newtonian material with a polymeric fluid
exciting from a die. The ratio of the diameter of the jet to the capillary diameter is
113% for the Newtonian case (low Reynolds number) whereas it can be as high as a
few 100% for the case of polymeric fluids. This is a very important phenomenon when
extruding products. The die-swell will need to be accounted for (think about how you
would make a square duct by extrusion). Again the die swell arises mainly because of
differences in the normal stresses along and perpendicular to the stream lines in the
tube. They relax as the fluid exists from the tube causing a contraction in the
longitudinal direction. In the above pictures the Reynolds number is 0.001. The fluid
on the top is a Newtonian oil, the second fluid is an elastic fluid prepared in the same
manner as discussed previously (From Boger and Walters, Rheological Phenomena
in focus (Elsevier 1993).

6
V. Entry flows : vortex enhancement

N1, e

In more complex flows, the material characteristics will strongly alter the flow fields.
The above pictures are examples of flow-visualization of a contraction flow at different
flow rates. As the flow rate is increased, the fluid behaves more and more as a elastic
material. The Deborah number, which compares the characteristic time of the process
to the characteristic relaxation time of the fluid, hence expresses how important
elasticity will be. As the Deborah number is increased, Vortices are enhanced. These
vortices can even become unstable and start to swirl in the die. Two often related
aspects are important in entrance flow, the normal stresses and the elongational
viscosity. The fluid used in the experiments was a solution of 0.04% High
Mw polyacrylamide in a water/corn syrup mixture. This fluid is a so-called Boger Fluid,
which has an essentially shear rate independent viscosity and a pronounced elasticity.
The Reynolds number in the above experiments were small, ranging from 0.0005 to
0.02. The Deborah Number was 0.08 and increased to 0.2. If the flow rates are
increased further, the vortices become unstable and dance in the barrel, causing
unstable flow. This is not related to turbulence, but is an instability due to the elastic
nature of the fluid.

The above pictures are from : D.V. Boger, D.U. Hur, R.J. Binnington, J. Non-Newt. Fluid
Mechanics, 20, 31 (1986) .

7
VI : Ductless Syphon

Tubeless syphon

McKinley et al (MIT)

D. Joseph and coworkers (UMinn)

When a liquid containing large macromolecules or high aspect ratio particles is


stretched the molecules and/or particles are aligned in the direction of stretching
resulting in a substantial increase of the resistance to flow. The elongational
viscosity can be many times higher as compared to the shear viscosity, especially at
high deformation rates. The above pictures show the open or ductless-syphon effect.
The slightest spilling will empty the beaker. When an initial thread of the material has
been drawn over the side the beaker empties itself. The material in the above
experiments is an aqueous solution f a water soluble polymer (polyox WSR301) in
water. (From Boger and Walters, Rheological Phenomena in focus,Elsevier ,1993).

8
VI : Reduction of Misting e

Elongational viscosity and elasticity of fluids can also be important in high Reynolds
applications, for example the break-up of a jet from a nozzle. The picture on the left
is for Water. The picture on the right is for a 200 ppm (!) solution of a high molecular
weight solution of a high Mw polyethylene oxide. (from JW Hoyt and JJ Taylor, Phys.
Fluis, 20 S253 (1977)). This is e.g. important in spraying or coating applications.
Another example of the same effect relates to drag reduction in turbulent flow.

9
Aim of this lecture:

1. Qualitative insight of the behavior of non-Newtonian fluids.

2. Discuss measurement techniques for a given problem

3. Demonstrate link rheological properties // microstructure

10
Rheology

Finite Deformation Tensors

We now have a way to determine the state of stress in a material. Now we need a
similar three-dimensional description of the state of deformation in a material. Relating
stresses and deformations leads us to the constitutive equations that we are looking
for. How do we proceed? Do we look to describe all the velocities in a fluid? This is of
not much use (fortunately)- we need the relative deformations.

At the end of this session we should have a description for:


- deformations (relative)
- deformation gradients

Lets keep this in mind:


If we write Hookes law in three dimensions, with the stress and strain being
proportional, the deformations tensors should be such that rotations do not give rise to
a stress. Likewise we will try to see if we can write Newtons law in three dimensions.

11
1. Deformation gradient tensor:

dx'
x
dx

z
Uniaxial elongation

dx = F.dx

Consider a cube of material as drawn above. Two points in that body are separated
by a vector dx at time t . Now let the body be deformed to a new state, e.g. the
deformation given above is a uniaxial elongation. The two points will move with the
body and the displacement vector between them will be stretched and rotated as
indicated by a new vector dx. Now we wish to relate dx (the present state) to dx (the
rest or past state at t). This will be given by another linear transformation, called the
deformation gradient tensor F.

Just as the stress tensor describes the force acting on any plane, the deformation
gradient tensor relates the state of deformation and rotation for any material point.

For the example of uniaxial elongation stresses will develop along the normal axis of a
piece of e.g. rubber. Keeping in mind Hookes law, we hence know what to expect for
the deformation tensor (which components will be non-zero?).

12
Simple shear ! = s/h
S

Solid body rotation "

Above are two other simple deformations which can be given to bodies. They are a bit
more critical to check whether or not we have a good deformation measure: The
simple shear deformation will lead to shear stresses, and again based on Hookes law
(proportional) now know to expect shear components in the deformation measure. By
the symmetric nature of the stress tensor, both the xy and yx components should
also be present in a suitable deformation tensor.

Likewise, a solid body rotation forms another critical test. Material elements are not
stretched, only rotated and no stresses should develop. A suitable deformation
measure should hence yield the identity tensor for this type of deformation

Now is F a good measure for deformations?

13
! (x (x (x $
# (x' (y ' (z'&
# (y (y (y & Deformation gradient tensor
F=#
(x' (y ' (z'&
## (z (z (z &&
" (x' (y ' (z'%

uniaxial extension Simple shear Solid body rotation


x = )x' x = x'+ 'y' x = x'! cos ) " y'! sin )
y = ) '1/ 2 y' y = y' y = x'! sin ) + y'! cos )
z = z' ' z = z' '
z = ) '1/ 2 z'
! 1 ' 0$ # cos ) " sin ) 0&
!) 0 0 $ F = # 0 1 0& F = %% sin ) cos ) 0((
F = # 0 ) '1/ 2 0 & # 0 0 1&
" % $ 0 0 1'
# '1/ 2 &
"0 0 ) %

If we look at what F evaluates to for a uniaxial elongation, we can be satisfied. When


we look at either simple shear or a solid body rotation, however, we see that we are
in trouble:

F is not always symmetric!

F is not zero for a solid body rotation!

F is hence not a good deformation measure for our purpose. It describes both rotation
as well as a shape change. Somehow we must eliminate this rotation. The material
response is determined only by the amount of stretching, not by a solid body rotation.

14
2. The Finger tensor:

Remove the rotation: multiply F by its transpose

Simple shear

" 1 ) 0% " 1 0 0% "$ 1 + ) ) 0%'


2
C-1 = F.FT
C!1 = $ 0 1 0' ( $ ) 1 0' = $ ) 1 0'
$ 0 0 1' $ '
# & # 0 0 1& $# 0 0 1'&

Uniaxial elongation

" *2 0 0 %
!1 $ !1 '
C =$ 0 * 0 '
$ 0 0 * !1'
# &

We know from matrix algebra that multiplying a matrix by its transpose, yields a
symmetric matrix. Can we, by doing this also remove the rotation and get a symmetric
tensor? We can express the deformation gradient F as a product of both a stretch and
a rotation. F= V.R. as a rotation followed by an inverse rotation gives I:
F.FT = V.R.RT.VT = V.VT
The dot product of F with its transpose is called the Finger Tensor, and it is a good
deformation measure. Note that other deformations measures are possible e.g. FT.F.
The physical meaning of the Finger tensor is that it relates the change in area within
the sample.

The good news is now that we can write Hookes law in 3D!
=GC-1

15
2. The Finger tensor:

Remove the rotation: multiply F by its transpose

Simple shear

" 1 ) 0% " 1 0 0% "$ 1 + ) ) 0%'


2
C-1 = F.FT
C!1 = $ 0 1 0' ( $ ) 1 0' = $ ) 1 0'
$ 0 0 1' $ '
# & # 0 0 1& $# 0 0 1'&

Uniaxial elongation

" *2 0 0 %
!1 $ !1 '
C =$ 0 * 0 '
$ 0 0 * !1'
# &

We know from matrix algebra that multiplying a matrix by its transpose, yields a
symmetric matrix. Can we, by doing this also remove the rotation and get a symmetric
tensor? We can express the deformation gradient F as a product of both a stretch and
a rotation. F= V.R. as a rotation followed by an inverse rotation gives I:
F.FT = V.R.RT.VT = V.VT
The dot product of F with its transpose is called the Finger Tensor, and it is a good
deformation measure. Note that other deformations measures are possible e.g. FT.F.
The physical meaning of the Finger tensor is that it relates the change in area within
the sample.

16
3. Neo-Hookean Solid:

Using the Finger tensor we can come up with a general form of Hookes law

Simple shear

$ 1 # 0' $ 1 0 0' $1 + # 2 # 0'


= G . C-1 ! = G.C = & 0 1 0 ) * & # 1 0 ) = G.& #
"1
1 0)
& ) & ) & )
% 0 0 1( % 0 0 1( &% 0 0 1 )(
! xy = G#

Uniaxial elongation

$+ 2 0 0 '
! = G.C = G.& 0 + "1 0 )
"1
& )
&% 0 0 + "1 )(
G
! xx = " p + G+ 2 = " G(1 + , )2
1+ ,
3, + 3, 2 + , 3
=G - 3, G = E,
1+ ,

The good news is now that we can write Hookes law in 3D!
=GC-1

Actually Hooke observed the proportionality between force and stress for a springy
body as early as 1678 and it took almost 150 years to formulate this model in 3D.

For an incompressible solid, the above calculation shows the link between the shear
modulus G and Youngs modulus E. For compressible materials, a Poisson-coefficient
needs to be used. For more details we refer to the textbook by Macosko
or the course Sterkteleer.

17
" )v x )v x )v x %
4. Velocity Gradient: $ )x )y )z '
$ )v y )v y )v y '
!v = $ '
$ )x )y )z '
)
$ z v ) v z ) v z'
# )x )y )z &

uniaxial extension

v x = *! x
(*!
vy = y
2
(*!
vz = z
2
" %
$ *! 0 0 '
$ (*! '
!v = $ 0 0 '
$ 2 '
$0 (*! '
0
# 2&

The velocity gradient tensor describes the steepness of the velocity variation as one
moves from one point to the other. This tensor has again two directions associated
with it : one of the velocity and one of the gradient. It operates on a local
displacement vector to to generate the magnitude and direction of the changes in the
velocity.

For the case of uniaxial elongation, it seems to give a good deformation measure.
But once again, solid body rotations and shear flows are more critical tests.

18
y
Simple shear U

v x = (!y d
vy = 0
vz = 0
x
no slip
" " 0 ( 0%
! U
!v = $ 0 0 0 '
linear velocity field: u x = ---- y
d
$ 0 0 0'
# &

stretching rotation

Lets consider the case of steady state shear flow. Again, we want the deformation
measure we use to yield a symmetric tensor in order to write our constitutive
equations. The kinematics of simple shear flow can be obtained by taking the time
derivatives of the displacement functions of simple shear we have used before.

We see that the velocity gradient tensor is not symmetric! Like the deformation
gradient F, the velocity gradient tensor contains both rotation as well as stretching.
A steady state shear flow can be viewed as a combination rotation and stretching. We
can remove the rotation from the velocity gradient by recalling that we could
decompose F in terms of a stretch and rotation
F= V.R if we differentiate this with respect to time:

F! = V
! !R + V !R
!

lim V = lim R = I
x'" x x'" x
lim F! = #v = V
! +R
!
x'" x

lim F! T = V
x'" x
! T
! +R ( )
! = 2D = (#v)T + #v
V

19
5. Rate of Deformation Tensor

T
2D = (!v) + (!v)

uniaxial extension

(" % " .! 0 +
1 *$ .! 0 0 0 .!
' + $ 0 !.! / 2 0 %' - = "$ 0 !.! / 2 0 %'
0 0
D= 0 !.! / 2 0
2 *$ ' $# 0 0
' $
!.! / 2& - # 0 0
'
!.! / 2&
)# 0 0 !.! / 2 & ,

Simple shear

! 1 $
# 0 -! 0&
' 0 -! 0$ 2
1 )!# ! 0 0 0$ * # 1 &
D= 0 0 0& + # -! 0 0& , = # -! 0 0&
2 )#" 0 0 0&% # &
( " 0 0 0% ,+ # 2 &
# 0 0 0&
# &
" %

Hence we obtained a new tensor 2D, which should be symmetric as is checked above
for the two types of deformation : uni-axial extension and shear flow.

As can be seen above, we again obtained a good measure to look at deformation


rates - it is symmetric in shear flow and you can check for yourselves that it yields I for
a solid body rotation.

20
General Newtonian Constitutive Equation

In the Newtonian model, the stress is assumed to be proportional to the


shear rate with viscosity being the proportionality constant.
Since the stress tensor is symmetric, it should be related to the rate of
strain, D, and not the velocity gradient tensor.

! = 2 (D y
U

" 1 %
$ 0 )! 0' d
$1 2 '
! = 2($ )! 0 0'
$2 ' no slip
x
$ 0 0 0' U
linear velocity field: u x = ---- y
$ ' d
# &

We now have derived all the tools that we need to extend Newtons law in three
dimensions. Note that the above equations relate the extra stress tensor and the rate
of deformation tensor. When we write the equation in its components for simple shear
flow, we find the definition that we know from e.g. transport phenomena or hydraulics.

The Newtonian constitutive equation is the simplest equation there is for viscous
liquids and forms the basis of fluid mechanics. When you are faced with a new
problem, this model will be the best place to start. Fluids which obey Newtons law are
e.g. low-molecular weight liquids, dispersions at small volume fractions of dispersed
phase, but even high Mw polymers at low deformation rates

21
uniaxial extension

" (! 0 0 %
D = $$ 0 ! (! / 2 0 ''
# 0 0 ! (! / 2&

+! = (!
) = 2 *D
)11 = !p + 2 *(!
) 22 = ) 33 = !p - *(!

) ! ) 22
*e = 11
(!
*e = 3 * TROUTON ratio

The above calculation shows the value of using the tensor form of Newtons law. The
result that the extensional viscosity is three times the shear viscosity is also observed
for polymer melts at low rates. It is actually often used to check the validity of results
of elongational rheometry and we will probably come back to this.

22
Linear Viscoelasticity

In this chapter we will see one of the most important aspects of rheology: the description of linear
viscoelasticity. Linear viscoelasticity is important for a number of reasons:
It is an important goal of rheology that we are able to describe and characterize material
behavior that is intermediate between solid and liquid behaviour.
The linear regime is important because it is the building block for further work: how
complicated the non-linear regime might be, we should always find the linear viscoelastic
behavior as a limiting case.
Last but not least linear viscoelasticity is of practical importance: For example for linear
polymer chains, measurements characterizing the linear viscoelastic properties provide insight in
to how the material will behave in some non-linear flows.

At the end of this chapter you should be able to:


define linear viscoelasticity.
understand it from an energetical point of view.
be able to describe linear viscoelasticity mathematically
know which properties can be determined experimentally and how they are interrelated :
relaxation modulus, compliance, storage and loss modulus and relaxation spectrum

23
What are viscoelastic fluids?

= fluids with a finite memory !

Memory = Memory = 0

Real fluid/solid

The memory point of view:

Viscoelastic fluids are defined as fluids which have a finite memory. The two material
classes we have worked with up to now; the perfect solid (Hookes law) and the
perfect liquid (Newtons law) do not obey this definition. A perfect solid will remember
the deformation indefinitely and return to its initial shape. A perfect liquid will stop
deforming instantaneously when the stress is removed and never recover to its
previous shape.

All real fluids and solids show intermediate behavior. I cannot recuperate the energy
that I put into a deformed solid. There will always be some internal dissipation of
energy. For example car tyres heat up during driving, not soo much due to the friction,
but due to the internal heat dissipation.

24
What are viscoelastic fluids?

= fluids that store and dissipate energy !

Energy storage Energy dissipation

Real fluid/solid

The energy point of view:

When we apply a stress and a body is deformed, this implies that workis being defined as force x
deformation, is done. The perfect solid will store this energy. According to Hookes law, the potential
energy
"
E = # d"
cannot be dissipated. If a perfect body would be deformed by a certain! xy xy
deformation g, and subsequently it
0 liquid on the other dissipates all the
would be released, it would oscillate indefinitely (in vacum). A perfect
energy , to keep the fluid going I need to put energy into the fluid.

All real fluids have a finite memory. Even water has a time constant of about 10-12 s. Deformed slower than
that it will flow, deformed more rapidly it will behave as a solid. Some materials have time constants on the
order of 1017s. Most polymeric liquids will have time scales which are on the order of 10-3 to a few seconds.
In addition , they will often show a wide distribution of relaxation times.

Whether or not you have to take viscoelasticity in to consideration when designing a process or product will
depend on the time scales of the product and the process.

25
1. Viscoelastic phenomena :
1.a. CREEP

Compliance

! ( t)
J( t) =
"0

The inverse of the creep experiment is an experiment were a strain is suddenly applied and the
resulting stress is measured. Either the a constant strain rate can be applied or a a constant
strain is suddenly applied. The stress in a viscous (Newtonian) liquid (VL) will drop to zero
instantaneously. The elastic solid (ES) reacts to a constant strain by an instantaneous stress
that remains constant in time.

What happens with a VE fluid in a stress relaxation experiment? As the name of the experiment
suggest, the stress gradually relaxes. We can convert this data into a RELAXATION modulus .
Note that the above examples are for shear deformations but the concepts and models we
discuss can be applied to other types of deformations. But because we are dealing with small
strains, typically only shear
deformations are of interest.

Questions :
Stress relaxation experiments are not always in the linear response regime. Think about what
condition needs to be satisfied for a step strain experiment to be a linear measurement.

26
1.B. stress relaxation

fast

Apply strain

Relaxation modulus

!( t )
G( t) =
"

The inverse of the creep experiment is an experiment were a strain is suddenly


applied and the resulting stress is measured. Either the a constant strain rate can be
applied or a a constant strain is suddenly applied. The stress in a viscous (Newtonian)
liquid (VL) will drop to zero instantaneously. The elastic solid (ES) reacts to a constant
strain by an instantaneous stress that remains constant in time.
What happens with a VE fluid in a stress relaxation experiment? As the name of the
experiment suggest, the stress gradually relaxes. We can convert this data into a
RELAXATION modulus . Note that the above examples are for shear deformations
but the concepts and models we discuss can be applied to other types of
deformations. But because we are dealing with small strains, typically only shear
deformations are of interest.

Questions :
Stress relaxation experiments are not always in the linear response regime. Think
about what condition needs to be satisfied for a step strain experiment to be a linear
measurement.

27
LDPE [Laun, Rheol. Acta, 17,1 (1978)]

The figure shown above contains the relaxation modulus for a low density
polyethylene at 150C (data from H.M. Laun, BASF, Rheol. Acta 1978). Data at two
different strains are shown. Note that the time scale over which the modulus
decreases expands up to a few 100 s!!! So this material has considerable memory for
the applied deformation. The data is represented on a log-log scale. The modulus
decreases rapidly, but it is not a single exponential decrease.

Remark:
Come back to this figure at the end of this chapter: then you will understand that the
line in the figures represents a fit to eight exponential relaxation times.

28
1.C. Sinusoidal oscillations

STRAIN:

! = ! 0 sin("t )

! = ! 0 exp(i"t )

STRAIN RATE

! = ! 0 " cos("t ) = ! 0 sin("t + 90 )

! = i "! 0 exp(i"t )

HOOKEAN SOLID

# = G! = G! 0 sin("t )
# = G! 0 exp(i"t )

NEWTONIAN FLUID

# = $!! = $! 0 " sin("t + 90 )


# = $i! 0 " exp(i"t ) = $i!"

The creep experiments are especially suited for the long time scales. The stress
relaxation experiments can access time scales between 0.1 (because of transducer
resolutions) and a few 100 s. So we need other types of experiments in which the
time scale can easily be varied. The small strain oscillatory measurements present
us such a tool!
The sample is deformed sinusoidal. After a few cycles of start-up, the stress will also
oscillate sinusoidal. For a Viscous Liquid (VL) and an Elastic Solid (ES) the response
can be easily calculated by invoking the appropriate constitutive equations. From the
above equations it can be seen that the stress in an ES will oscillate IN PHASE with
the applied deformation, whereas for a VL the stress will be DEPHASED by 90 with
respect to the oscillatory deformation. Again, we found a way to compare a elastic
solid and a viscous liquid. Note also that for the VL the stress amplitude will depend
on frequency whereas for an ES it will not!

29
VISCOELASTIC MATERIAL

! = ! 0 sin("t + # )

! = ! 0 exp(i"t + # )

complex mod ulus:

! = G &$

! = G & $ 0 sin("t + # )

! = G & $ 0 [sin("t ).cos(# ) + cos("t ).sin(# )]


( ) ( )
! = G & cos(# ) % $ 0 sin("t ) + G & sin(# ) % $ 0 cos("t )

! = [G '% sin("t ) + G "% cos("t )]$ 0

! = (G '+ iG " )$

G' '
Storage and Loss modulus tan ! =
G'

Now what do we get for a VE liquid: The total stress can be viewed as the combination of in
part a viscous and in part an elastic component, resulting in a complex modulus with a real
elastic part, and an imaginary (viscous) part The stress response will be dephased by an angle
. This angle will be a measure for how elastic or how viscous my fluid is. If the fluid is viscous
the phase angle will tend to 90. When the fluid is dominantly elastic, the phase angle will tend
to zero. The phase angle is hence a nice way to characterize the material on a viscoelasticity
scale. It is convenient to analyze the stress wave into two waves, one in phase and the other
one out of phase with the strain wave. This decomposition suggest the existence of two
dynamic moduli:
G the in-phase, elastic or storage modulus
G the out-of-phase, viscous or loss modulus.
Hence we see the definition of VE in terms of energetic behaviour reappearing from this
experiment. The figure on the right shows oscillatory data on the same LDPE as shown before
in a creep experiment. We will try later to calculate the creep from the dynamic moduli (a nice
feature of linearity all Linear viscoelastic functions are interrelated)

30
Newtonian MATERIAL

# = %"!
# = [G '! sin($t ) + G"! cos($t )]" 0
# = ( G" ) "
G" =
G = Loss modulus tan ! = ?
!=?
Elastic " (Hookean) MATERIAL

$ = G#
$ = [G '! sin(%t ) + G"! cos(%t )]# 0
$ = (G ' ) # tan ! = ?
G' =
!=?

For a Newtonian liquid:


G is known as the viscous loss modulus. It characterizes the viscous contribution to
the stress response. What is the phase angle in this case?

For an Elastic material :

For Hookean solids, stress is proportional to strain and we find another limiting result
G is thus known as the elastic storage modulus. It characterizes the elastic
contribution to the stress response. Remind yourseld : what is the phase angle in this
case?

31
VISCOELASTIC MATERIAL
! = ! 0 exp(i"t + # )

complex vis cos ity :

! = $3 %!

! = ( $'+ i$" )i"% = ( $"& i$')"%

G"
$' =
"
G'
$" =
"
1/ 2
-' G '* 2 ' G "* 2 0 1 3
$3 = /) , + ) , 2 = G
/.( " + ("+ 2
1 "

Another way to view the experiments is in terms of of a sinusoidal strain rate. This
leads to the definition of a dynamic viscosity. This is also a material function. We can
now define a dynamic viscosity for the magnitude of the viscous stress to the
strain rate and an elastic part of the dynamic viscosity .

The typical low frequency behaviour for the materials function for polymer melts and
solutions is sketched in the above graph (A). It shows a plateau modulus G0 and
decreases with 2 for G and proportional to for G at low frequencies. This is an
important feature to check when doing measurements. For rubbers (fig: B) on the
contrary G is essentially constant over the entire frequency range.

32
Mirror
relation

Terminal zone :

Time (s)

When the frequency range for a polymer melt is extended, we find again the
behaviour that is known from the courses on polymer science. The different regions
are indicated on the graph.

The graphs on the right hand side also indicate the link between dynamic moduli
and the relaxation modulus. In an oscillatory experiment, the sample is
interogated at each frequency. The relaxation modulus corresponds to an integral
over the spectrum of relaxation times : i.e. at short time all modes contribute to the
stress, whereas some start to relax as time increases.

For a VE liquid the limiting relations for the terminal zone, shown above should
always be true when the material is tested in the linear viscoelastic regime. Hence
such relations are extremely useful for testing the internal consitency of data. You can
use the relations
above to check whether data from different types of experiments agree and even if
data from non-linear experiments gives in the limit for vanishing shear rates the VE
result.

33
Effect molecular weight on
the LV moduli of polymer
melts

The above graphs show the dependence of the moduli on molecular weight for a
set of monodisperse PS polymer melts. These results point us to an application of
rheology as an analytical tool : LV measurements will give us detailed information
about the distribution of molecular weigths. To do this well need to develop a
theoretical description. Even than theres a catch : the calculation of the molecular
weight distribution is an example of an inverse problem (and hence intrinsically and
often numerically difficult).

34
Effect of volume fraction
on the dynamic moduli of
dispersions ! d 2* $
# 2&
" dr % 5+ ! G ') 6, c $
F( r ) = = # ( &
d c ' kT 0.639 N " kT +d 3c %
104 "c

0.421
10 3
0.426
: pairwise interaction potential
0.437 N : coordination number (7.5)
0.452
102 0.455
G: high frequency modulus
0.471 dc : particle core diameter
101
0.482
0.502
c : core volume fraction
0.515
0.534
100 !(r)/kT !d/kT
G' (Pa) G"(Pa)

-1
10

10-2
10-2 10-1 100 101 102

! (rad/s)

!(r)/kT

The above graphs show the dependence of the storage modulus on frequency and
volume fraction. Again an application of the linear viscoelastic measurements arises :
we will be able to use rheological measuremenbts to characterize the interactions
between particles. As we will see in the chapter on microrheology of dispersions, we
will make use of rheology as a sensitive measurement technique to asses dispersion
stability and predict the (non-linear) viscosity curves.

35
2. Differential models : The Maxwell model

Phenomenological models

Hookean Spring

! 1 = G" 1
Newtonian Dashpot

! 2 = " $ #! 2
Maxwell Kelvin-Voight

Lets now try to describe the dual nature of these fluids and solids by simple
mathematical models. We can start from simple phenomenological tools. A Hookean
spring will be our model for a perfect solid. A dashpot will be our conceptual model for
a perfect fluid. We can now start combining these elements in parallel or in series and
so on. We could as well draw an analogy on electrical elements, using resistors and
capacitances.

Maxwell (1867) first proposed the in series model for the viscosity of gases.

Question : from the above cartoons, which one will be the best for describing a fluid?
Which one will describe a solid?

36
Maxwell model = VE Fluid ! = !1 + ! 2

two elements in series " = "1 = " 2


! = !1 + ! 2

"! 1 " 2
!= +
G0 #0

%# (
" + ' 0 * "! = #0 !
& G0 )

" + $"! = #0 ! = relaxation time

Now lets assume we take the Maxwell model, using the elements in series. When we
apply a deformation to this model the total deformation will be split by both elements.
The stress in both elements is the same and equal to the total stress. Assuming a
homgeneous deformation, the deformation rate can also be assumed to consist of a
sum of the two individual motions.

We can then rewrite the equations, leaving us with a differential equation. This
differential equation is know as the MAXWELL model; In the next few pages we will
try to see if this model can capture some of the observed phenomena of linear
viscoelasticity.

37
Relaxation (Step strain) experiments for a Maxwell element
12
G0

! = !0 t $ 0 10

d" "

"+# =0
6

G(t)
dt 4

t = 0; " = G0 ! 0 2

"(t ) 0

= G(t ) = G0 % exp( &t / #) 0.01 0.1 1 10

! Time (s)
! [rad/s]

Fading memory

Lets first calculate the response of the Maxwell model to a step in strain. As boundary
conditions we impose the strain of g0 at times > 0. We can easily calculate the stress
at time zero (purely elastic response) and solving the differential equation gives use
the time evolution of the stress and of the relaxation modulus.

The full line on the curve gives the predictions of the above equations for a G0 of 10
Pa and a relaxation time of 1 s on a semi-logarithm plot. In reality, a single
exponential is not always observed (remember the relaxation curve for the LDPE we
showed previously). The dotted line shows a more gradual evolution of the relaxation
modulus. We will see how we can modify the Maxwell model to take into account this
behaviour.

38
Dynamic moduli

10 G0

G0 ! 2 " 2
G '(! ) =
1 !

2 2
1+ ! "
0.1

G0 !"
G "(! ) =
Moduli [Pa]

1 + ! 2"2 0.01
G'
G"

0.001

0.01 0.1 1 10 100

w [rad/s]

Likewise you can calculate the response of a Maxwell model to an oscillatory strain. It
is most easily done by using complex notation.

What you can easily see from the evolution of the moduli is that for frequencies
smaller than the inverse of the relaxation time, a dominantly viscous response is
observed. This can be extended to all processes with slow motions indeed the
dashpot or Newtonian behavior will dominate. At high frequencies the elastic
behavior dominates. For rapidly changing stresses in general, the derivative term in
the Maxwell model equation dominates and at short times this model then approaches
elastic behavior.

39
Do Maxwell Fluids exist?

Water oil emulsion Concentrated surfactant solution

Do Maxwell fluids exist? Well at least some fluids do behave according to the simple
maxwell model, at least to a good approximation. These are fluids in which one very
dominant relaxation mechanism exists. Two examples are given above:

Emulsions : here the VE comes from the interface, the material now can store
elastic energy as the droplets, which have an interfacial tension, can deform. The
relaxation time of these fluids is determined by the medium viscosity, the droplet size
and the value of the interfacial tension. When the droplets are fairly uniform in size,
the system will respond with essentially one single relation time. The graph on the left
compares two emulsions which were subjected to different mixing histories.

Surfactants : A lot of shampoos come relatively close to being Maxwell fluids! Such
surfactants can form - in a certain range of concentration - rodlike micellar objects.
These objects are sometimes called Living Polymers as they resemble polymeric
fluids but they can break and reform without too much effort. The characteristic time
scale of these systems is the breaking or reconnection time.

40
Can we describe polymer melts?

As we already indicated a single exponential is not always observed. The relaxation


curve for a PDMS sample is shown above and the LDPE was shown previously. We
will see now we can modify the Maxwell model to take into account this behavior.

Come back to this page and explain what the right hand side graph means.

41
Generalized Maxwell model

!TOT = % ! i
i

! i + " i !! i = #i $!

G(t ) = % G0 i exp( &t /" i )


i

Jeffreys model

% d$! (
! + "1!! = #' $! + " 2 *
& dt )

An obvious way to generalize the Maxwell model is to allow for multiple relaxation
times. This can be done by considering multiple dashpot-spring combinations in
parallel. In this manner, several relaxation times are available to fit our experimental
data.

The relaxation data of the PDMS on the previous page could be described by
selecting a set of relaxation times equidistantly spaced on a logarithmic axis and
determining the Gks for the different modes by a linear regression analysis. (k = 0.01,
0.1, 1, 10, 100 s, Gk=2*105 , 105, 104,102,10) .

Some molecular theories, like the Rouse theory for polymer dynamics (see later)
suggest an infinite series of relaxation times where only two fit parameters remain.
"k
Gk = G0
! "k
k
"O
"k =
k2

42
3. General viscoelastic Integral model

&(t ) = $ G(t ! t ' )'%(t ' )

t
&(t ) = # G(t ! t ' )d%(t ' )
!"

t
d%
&(t ) = # G(t ! t ' ) dt '
!"
dt '

fluid
t
&(t ) = G(0)%(t ) ! G(")%( !" ) ! # %(t ' )dG(t ! t ' )
!"
Reference state @ t=0

&(t ) = # M (s )%(s#)ds s = t ! t'


! (t) = $ M(s)" (s)ds
0

s =t !t'
dG(s )
M (s ) = memory function =
ds
dg(s)
M(s) = is the memory function
ds

An arbitrary evolution of deformation in time can be approximated by a series of


stepwise changes . If we assume that the stress a time t resulting from the given
motion equals the sum of the partial stresses at time t from the individual strains
at time (t) (this is what is called the linearity or Boltzman principle), the total stress
can be calculated as indicated above.

When the steps are taken to be infinitesimally small, the summation becomes an
integration. Note that in rheology it is customary to use the configuration at time t as
the reference. The equation can then be transformed by an integration by parts.
Noting that G()=0 for fluids and is finite at time t = that (t)=0 is a reference for
the deformation.

43
What is the relaxation function?
For a simple Maxwell model we used a single exponential

G(t ) = G0 exp( !t / +)

For a generalized Maxwell fluid this becomes

G(t ) = " Gi exp( !t / + i )

We can replace the discrete relaxation times by a continuous spectrum

*
# !t &
G(t ) = ) F ( +) exp% ( d+
0 $ +'
Often one use a logaritmic scale : the relaxation spectrum is defined as follows

*
# !t & d+
G(t ) = ) H ( +) exp% (
0 $ +' +
Powerful way to characterize
materials - albeit in the Linear region

The use of G(s) does not lead directly to a specific equation for the stress relaxation
function. Specific expressions can be developed. By using a simplest expression is a
single exponential decrease. In this manner we obtain the integral formulation of the
Maxwell model. The integral form of the generalized Maxwell model can be easily
obtained by using the relaxation spectrum as defined above.

By using a relaxation spectrum, any arbitrary relaxation function can be fitted


accurately in a unique manner. This provides a very powerful tool to characterize the
rheological behavior of polymers. The relaxation spectrum allows one to obtain a
dynamic fingerprint of the behavior of polymers. For linear polymers, the relaxation
spectrum can be linked to the molecular weight and the molecular weight distribution.

44
Oscillatory flow G and G from the general linear viscoelastic fluid:

! = ! 0 " cos("t )
t
# = ' G(t $ t ') % ! 0 " cos("t ')dt
$&

&
= ! 0 " ' G(s ) % cos("t $ "s )ds
0

& &
= ! 0 " ' G(s ) % cos("s )ds cos("t ) $ ! 0 " ' G(s ) % sin("s )ds sin("t )
0 0

&
G " = " ' G(s ) % cos("s )ds
0
&
G ' = " ' G(s ) % sin("s )ds
0

For all linear viscoelastic materials, the shear stress is expected to be a sinusoidal
function when a sinusoidal strain is applied. The above calculation demonstrated how
the response for an arbitrary linear viscoelastic fluid can be calculated (continued on
the next page). The only mathematical tool we use is that
cos(a+b)=cos a.cos b + sin a . sin b

45
( ( N
" %
G " = , * G(s ) ! cos(,s )ds = , * $ + Gi e ) - i s ' ! cos(,s )ds
0 0 # i =1 &
N ,- i
= + Gi
i =1 1 + (,- i ) 2

(
G ' = , * G(s ) ! sin(,s )ds
0

N (,- i ) 2
= + Gi
i =1 1 + (,- i ) 2

Weve assumed the relaxation spectrum to consist of a discrete spectrum of


relaxation times. The equations that we have obtained were actually used to describe
the moduli of the LDPE by Laun [Laun, Rheol. Acta, 17,1 (1978)]

46
Oscillatory flow !'(" ) = ? Complex viscosity
G"
!'(" ) =
"
1,
= . G(s ) $ " $ cos("s )ds
"0

1 , %, H ( # ) - # (
s
= . '. $ e d# * $ " $ cos("s )ds
" 0 '0 # *)
&

H ( # ) %, (
, s
-
= ' . $e # cos("s )ds * $ d#
.
0 # '
&0 *)

%, ( b
+ ' . $e - bt cos(at )dt * = 2 2
for b > 0
&0 ) a +b

,
H ( #)
!'(" ) = . d#
0 1 + ("#) 2

Here we give another example how a VE function, in this case the complex viscosity -
can be calculated. This time we have used a spectrum rather than a discrete set. You
can try to show the what the equation becomes for ()

47
Some relations between viscoelastic functions

2 # G'
G( t) = % ' ! sin('t)d'
&0'

#
( = % H( )))d ln )
$#

2 # G'(' )
(= % d ln '
& $# '

dG' 1 d2 G'
H( )) " +
d ln(' ) 2 d(ln ' )2 1/ ' = ) / 2

The relaxation spectrum also gives us a tool to interrelate the different viscoelastic
response functions. The different response functions can hence be calculated from
each other. This can be either time saving or can be used to verify the internal
consistency of experimental data. Combining data from different types of experiments
can also be used to calculate the relaxation spectrum (it is an ill-defined mathematical
problem).

48
Conclusions
Defined and described linear viscoelasticity.
Using rheology as a spectroscopic technique!

MATERIAL FUNCTIONS :

Our goal is to obtain constitutive relations for our fluids -


as this is too difficult for a complex flow - we try to find a
number of functions in standard flows - the material
functions relating to linear viscoelasticity are

relaxation modulus G(t)


Compliance J(t)
storage and loss modulus [G() and G()]

For the case of linear VE they are interrelated by the


relaxation spectrum H()

This chapter we have taken a big step in describing rheological phenomena. Linear
viscoelasticity is important in practical applications. Some real life deformations will be
in the linear regime (especially for solid state rheology). Secondly, understanding and
describing the fading memory effect is an important aspect of rheology. We have also
provided a number of material functions that can be used to calculate relaxation
spectrum and that are interrelated. In the next chapter we will now see how these
material functions can be made non-linear. Then we will learn how all these
properties can be measured.

49
Applied Rheology

Generalized Newtonian
Fluids

Although Newtons law is of much practical use, at the beginning of this century
researchers started to realize that not all fluids obey this simple linear relation. Many
colloidal suspensions and polymer solutions show a decrease of the viscosity as the
shear rate is increased.

In the following, we will try to extend Newtons law in the most simple fashion, by
making the stress depend on the rate of deformation (or better : on an invariant of the
rate of deformation tensor). But as we are talking about viscosities and rate of
deformations, maybe it is useful to start of with indicating a few orders of magnitudes
which are likely to be encountered

50
Liquid [Pa.s]

Molten glass 1012


Asphalt 108
Molten Polymers 103-105
Syrup 101
Paints 100-101
Glycerin 100
Light Oil 10-1
Water 10-3
Air 10-5

The SI unit of viscosity is Pa.s (Pascal.second). The cgs unit of Poise (P) is also often
used. 1 poise = 0.1 Pa.s. Centipoises are also often encountered, as 1 cP is the
viscosity of water.

51
Process shear rate [s-1] Applications

Sedimentation 10-6 -104 Medicines, Paints,...


Leveling 10-3-10-1 Paints, inks
Draining under gravity 10-1 Emptying tanks
Extrusion 100-102 Polymer melts, Dough
Chewing, Swallowing 101 -102 Food
Dip Coating 101 -102 Paints, confectionery
Mixing and stirring 101-103 Manufacturing liquids
Pipe flow 100-103 Pumping, blood flow
Spraying, brushing 103 -104 atomization, painting, spray drying
Rubbing 104 -105 Skin cream and Lotion
Injection molding 102 -105 Polymer melts
MILLING 103-105 Paints, inks, coatings
coating flows 105 -106 paper
Lubrication 103-107 Engines

The range of shear rates encountered during various processes spans many
decades. Rheological properties are important in slow processes such as
sedimentation; some products will be tailored too have a very high viscosity at low
shear rates for a good shelf-life. Some processes such as rubbing or chewing
might seem odd to you, but the rheological properties determine to a large extent the
waqy these products feel . Coating flows and printing operate at high rates. The
flows, however, are still laminar.

It will be quite a challenge to develop measurement techniques that can embrace this
entire shear rate range!

52
Do real fluids behave according to Newtons law?

From Barnes (1989)

Just some examples of Non-Newtonian fluids are given above. Human blood (a) is a
complex fluid with a very variable rheological behaviour. The most important
consituents (rheologically speaking) are the red blood cells, which implies that it can
be viewed as a dispersion of deformable particles. Another complex dispersion is
yogurt (rheology can be used here as a tool to detect fermentation). Another natural
material with a pronounced shear-thinning behaviour is Xanthan gum (b) . Here the
structural units are macromolecules which interconnect to form a network that is
constructively broken down by the shear flow.

But increasing the shear rate does not always result in a decrease of the viscosity. In
some cases a shear-thickening behaviour is observed, e.g. for concentrated
dispersions containing large particles or associating polymers. As an examoke data
for latex spheres in diethyleneglycol is given in fig (d).

Note that a nonlinear behaviour of the viscosity not necessarilly implies visco elastic
behaviour.

53
1. General viscous fluid

! = f(2D)

! = " pI + a1 # D + a2 # D2 + a3 # D3 +...

A. Cayley Hamilton theorem

! = " pI + f1 # 2D + f2 # (2D)2 Reiner-Rivlin CE

f1(I2D , II2D , III2D )


f2 (I2D , II2D , III2D )

Scalar functions of the Invariants of 2D

To describe non-linear properties the governing equations have to made non-linear.


Newtons law stated a proporionality between stress and rate of deformation (y=
A1+A2.x ). To make this equation nonlinear we can add higher order terms in a power
series (y= A1+A2.x +A3.x2 +)

The simplest manner to extend Newtons equation is to have a stress tensor that only
depends on the rate of deformation tensor. The general expression can be expanded
in a power series. The isotropic term is of course the pressure term.

The Cayley-Hamilton theorem helps us in reducing the number of terms. As we are


working in a two dimensional space, all terms of higher order then 2 will be zero. The
functions f1 and f2 are scalar functions of the invariants. This equation is known as the
REINER-RIVLIN constitutive equation. All that remains to be done is to determine f1
and f2

54
1. General viscous fluid

B. Do we need all invariants of D? e.g. et us consider shear flow

! '! 2 0 0$ I2D = tr 2D = 0
( )
!0 '! 0$ # & 1
2D = # '! 0 0& (2D) = # 0 '! 2 0&
2 II2D = I2D ! tr (2D2 )
# & 2
"0 0 0% ## 0 0 0&& IIID = det(2D) = 0
" %

C. Do we need the second order term in 2D?

$ "! 2 0 0'
$ 0 "! 0' & )
! = # pI + f1& "! 0 0) + f2 & 0 "! 2 0)
& ) && 0 0 0))
% 0 0 0(
% (

! xx # ! yy = 0
! yy # ! zz = f2 "! 2

Firstly : The invariants of 2D can be easily computed. As it turns out, we do not need
all invariants, in simple shear flow two of the three will be zero.

Secondly: we can reduce the equations even more: the second order term gives rise
to normal stresses that are however not physically observed. No N1 and a positive
N2 think about the rod climbing experiment this is not what we observed. Typically
N1 = xx-yy will be positive and N2 = -0.1 N1

We end up with a general equation for the GENERAL VISCOUS FLUID:

# = "pI + !(II2 D )2D

55
1. General viscous fluid

! = " pI + f1(II2D ) # 2D

The viscosity is now replaced by a function of the second invariant of 2D

for a shear flow this becomes:

()
! xy = "1 #! 2 $ #!

and different forms for this function have been proposed

So we have gone through a lot of trouble to prove that the most sensible and
mathematically and physically correct thing to do is to replace the viscosity by a
scalar function that depends on the invariant of the rate of deformation tensor. For a
simple shear flow, the viscosity will depend on the magnitude of the shear rate.
Several forms have been proposed for this function. We will go through the different
models, distinguishing them by the number of parameters that are used.

56
Viscosity curves :

Styrene-ethylacrylate in H2O
102 # = 0.50
2a = 250 nm Data from Laun (1984)
101 0.47

100
0.43
(Pas)

10-1
"

10-2 0.34
0.28
0.18
10-3 H2O

10-3 10-2 10-1 100 101 102 103 104 105

! (Pa)

The graph above shows the viscosity versus shear stress for aqueous suspensions
of charged polystyrene-ethylacrylate spheres with a particle radius of 125nm. The
volume fraction was varied from 0 to 0.50, temperature was kept constant at 25C.
This graph reveals the typical fetaures that suspensions can display : At low volume
fractions the viscosity of the suspensions remains essentially Newtonian and is only
slightly greater than that of the dispersion medium. At higher volume fractions, the
occurrence of shear-thinning and shear-thickening is very obvious in this figure. At
very high volume fractions and low shear stresses, the Newtonian plateau tends to
disappear, and a yield stress starts to develop.

Not visible in the figure are eventual time effects. Especially for aggregated systems,
gradual time-dependent changes in the viscosity can be observed, i;e. thixotropy is a
typical property of such systems.

The visco-elastic nature of suspensions will also be addressed in this chapter.


Interaction forces between colloidal particles are potential forces and are elastic in
nature. They will, for example, show up in the oscillatory testing of dispersions. Normal
force differences are not very important.

57
Spherical particles : Zero shear viscosity

Dilute case (<0.05) (Einstein, 1906,1911)

" = "m [1+ 2.5!]


Semi-Dilute case (<0.10) (Batchelor, 1977)

[
" = "m 1 + 2.5! + 6.2!2 ]
Up to maximum packing (Krieger-Dougerthy,1959)
)[ ( ]' max
0 ' -+ & ( ) (m #
( = (m .1 ) [(] = lim $$ !
. 'max + '*0% '(m !"
/ ,

Einstein calculated the viscosity dissipation produced by the flow around a single sphere. The result is given above. It is only valid for low
concentrations, when the suspensions is dilute enough that the flow field around one particle is not appreciably influenced by the presence of
neighbouring spheres. The equation states that the size does not affect the viscosity. This is true as long as the matrix cn be treated as a
continuum

When two spheres are close enough so that the drag on one particle is affected by the drag on a second one, the spheres are said to interact
hydrodynamically. The velocity field around one sphere is felt by the another one. Such interactions which are far more complicated to
calculate. Interactions between two spheres lead to a term that is proportional in 2, three body interactions lead to a term proportional to 3 and
so on. Following an idea by Arrhenius, we can treat the suspension as an effective medium. The addition of a particle leads to an incremental
increase of the viscosity. Ball and Richmond (1980) suggested that this viscosity increments is proportional to :

Where [] is the intrinsic viscosity, which is the dilute limit of the viscosity increment per unit volume fraction, divided by the particle volume
fraction.

Intergration of this equation leads to the Krieger Dougherty equation (1959), which is an emperical equation for suspensions of spherical and
non-spherical shape. Often the product of []max is constant and approximately equal to 2.

[)])md(
d) =
&1 ' ( #
$ (m !"
%

" # "m
["] = lim
!$0 !"m

58
Spherical particles : Zero shear viscosity

1000
max
Einstein
Batchelor
Krieger-Dougherty
100
s
"/"

10

0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

This figure compares the predictions for the Einstein, Batchelor and Krieger-
Dougherty equation. For the KD equation, the divergence of the viscosity as the
volume fraction approaches maximum packing is observed. For hard spheres,
randomly packed, m ~0.63-0.64 (which will be relevant for the low shear viscosity).

59
Spherical particles : Zero shear viscosity

Particle size distribution effects


Volume fraction dependences of latices

This figure shows some data on latices : Data on PS spheres


x 5 mm (Manley and Mason, 1954)
PS .42, 0.87 mm (Saunders, 1961)
low shear viscosity, High shear viscosity (Krieger 1972)

These equations only pertain to suspensions of monodispersed particles. For high


volume fractions, effects of particle size distribution, and particle roughness are
observed. The figure on the right shows the evolution of the relative viscosity against
the fraction of large particles for a bimodal dispersion with 5:1 particle sizes, at various
total volume percentages. The arrow P-Q shows the reduction of the viscosity by a
factor of 50 when the volume fraction is kept constant, but the suspensions are
replaced by a 50-50% mixture of large and small spheres. The arrow P-S shows that
the volume fraction can be increased from 60 to 75 without increasing the viscosity
(figure From Barnes et al, 1989). These issues are of extreme importance for slurry
transport ot the design of products.

60
Viscosity curves :

Styrene-ethylacrylate in H2O
102 # = 0.50
2a = 250 nm Data from Laun (1984)
101 0.47

100
0.43
(Pas)

10-1
"

10-2 0.34
0.28
0.18
10-3 H2O

10-3 10-2 10-1 100 101 102 103 104 105

! (Pa)

The graph above shows the viscosity versus shear stress for aqueous suspensions
of charged polystyrene-ethylacrylate spheres with a particle radius of 125nm. The
volume fraction was varied from 0 to 0.50, temperature was kept constant at 25C.
This graph reveals the typical fetaures that suspensions can display : At low volume
fractions the viscosity of the suspensions remains essentially Newtonian and is only
slightly greater than that of the dispersion medium. At higher volume fractions, the
occurrence of shear-thinning and shear-thickening is very obvious in this figure. At
very high volume fractions and low shear stresses, the Newtonian plateau tends to
disappear, and a yield stress starts to develop.

Not visible in the figure are eventual time effects. Especially for aggregated systems,
gradual time-dependent changes in the viscosity can be observed, i;e. thixotropy is a
typical property of such systems.

The visco-elastic nature of suspensions will also be addressed in this chapter.


Interaction forces between colloidal particles are potential forces and are elastic in
nature. They will, for example, show up in the oscillatory testing of dispersions. Normal
force differences are not very important.

61
1. Large particles: Hydrodynamic Forces

Stokesian dynamics simulations (J. Brady and coworkers)

So far, we have only discussed the effects of particle volume fraction on the increase of the low shear viscosity. The viscosity of suspensions is,
however, strongly dependent on shear rate. The shear rate dependence of the viscosity occurs because shear flow will be able to distort the
equilibrium microstructure. The interparticle spacings will be altered, changing the magnitudes of both the hydrodynamic and the interparticle
forces. This will lead to shear thinning and shear thickening.

The microstructural changes in flowing dispersion can be complex in nature. The mechanism of shear-thinning in hard sphere suspensions
has been elucidated by computer simulations. J. Brady and coworkers introduced a technique called Stokesian Dynamics (see e.g. Brady
and Bossis, Annual Review of Fluid Mechanics Vol. 20: 111-157 (1988)), where the Stokes equations are solved for all particles. The distortion
of the equilibrium structure changes the interparticle spacing, and the Brownian contribution to the viscosity changes. Particles can be even
observed to line-up into strings parallel to the flow direction. When the shear rate is increased, particles can, however, be brought into close
proximity by the flow leading to the formation of so-called hydroclusters. The deformation of such clusters is difficult as the small separation in
between the particles lead to large lubrication stresses in the gap. For hard sphere systems, we will see that this leads to shear-thickening.
When colloidal forces are important, the microstructural reorganizations can have an even more complex dependence on the flow rate.

We will now first address the question if we can predict the transition to shear thinning.

62
Scaling for Brownian hard spheres

> 0.3 - shear rate dependence of the viscosity


Brownian motion is suppressed if

! $1 < t d
a2 k T
td = = a 2/ B
D0 6#"m a
" a ?!
Pe = m
k BT
As increases:
replace medium by an effective viscosity

Dimensionless shear stress

"susp ( #! ) $ #! $ a 3
!r =
kT

Distortion of the particulate microstructure will only occur when the rate at which the
particle equilibrium is recaptured is smaller than the rate at which it is being deformed.
The rate at which the structure regains its configuration is controlled by the particle
diffusivity, D0. For spherical particles the diffusivity is given by the Stokes-Einstein
relation: kBT
D0 =
6!"ma

Where kB is Boltzmanns constant, T is the temperature, m is the medium viscosity


and a is the particle radius. The time tD for a particle to diffuse a distance equal to its
radius is given by a2 / D0. The Pclet number is a dimensionless group which will
compare the time scales for diffusion and convection, or weigh the hydrodyanmic
forces with respect to the Brownian forces. When is the Pclet number small?

The definition of the Pclet number implies that the particles are seeing the viscosity
of the medium. Especially at higher volume fractions, it is better to use the effective,
suspension viscosity as the medium viscosity that a particle feels. Hence a
dimensionless shear stress is more appropriate.

63
(
(r = (susp
med
= f &$% ',)r ,t ,Pe,Re #!"
(r = f &$% ',Pe #!"

10000

0r = f .,- 1, / +)*
"r0
"r#
max!
1000

'[0]1max
& 1 #
0r = $1 ' ! r
100

% 1max "
"

10

& 0 ' 0r ( #
0r = 0r ( + $ r0 !
% 1 + b/r " 1
0.0 0.2 0.4 0.6 0.8 1.0

!
& 0 ' 0r ( #
0r = 0r ( + $ r 0 m!
% 1 + (b/r ) "

64
Shear thickening

16

14

12

10

r
!
6

4 Hydrodynamic viscosity
Brownian viscosity
Total viscosity
2

1e-2 1e-1 1e+0 1e+1 1e+2 1e+3 1e+4 1e+5

Pe

Stokesian dynamics simulations (J. Brady and coworkers)

So far, we have only discussed the effects of particle volume fraction on the increase of the low shear viscosity. The viscosity of suspensions is,
however, strongly dependent on shear rate. The shear rate dependence of the viscosity occurs because shear flow will be able to distort the
equilibrium microstructure. The interparticle spacings will be altered, changing the magnitudes of both the hydrodynamic and the interparticle
forces. This will lead to shear thinning and shear thickening.

The microstructural changes in flowing dispersion can be complex in nature. The mechanism of shear-thinning in hard sphere suspensions
has been elucidated by computer simulations. J. Brady and coworkers introduced a technique called Stokesian Dynamics (see e.g. Brady
and Bossis, Annual Review of Fluid Mechanics Vol. 20: 111-157 (1988)), where the Stokes equations are solved for all particles. The distortion
of the equilibrium structure changes the interparticle spacing, and the Brownian contribution to the viscosity changes. Particles can be even
observed to line-up into strings parallel to the flow direction. When the shear rate is increased, particles can, however, be brought into close
proximity by the flow leading to the formation of so-called hydroclusters. The deformation of such clusters is difficult as the small separation in
between the particles lead to large lubrication stresses in the gap. For hard sphere systems, we will see that this leads to shear-thickening.
When colloidal forces are important, the microstructural reorganizations can have an even more complex dependence on the flow rate.

We will now first address the question if we can predict the transition to shear thinning.

65
Non-Linear Viscoelasticity

Phenomena and modeling

In the previous two chapter we have gradually added complexity to the description of material behaviour.
Still the cases presented before are only limiting cases of material behaviour. The generalized Newtonian
fluid describes non-linear but inelastic materials at large deformations. The Linear Viscoelastic fluid
presents a viscoelastic material but it will only be valid for small deformations. Hence two parameters are
important, one being the magnitude of the deformation e.g. strain amplitude, the other being the
Deborah number which is defined as the characteristic relaxation time divided by the characteristic time of
the process.

#
De =
"! !1

???
0 N E

LVE
We now have to start describing a very wide range of nonlinear phenomena. Let us first try to make a round
up of the phenomena we at least want to describe
De

66
1. Non-linear VE phenomena
2. Additional mathematical tools
3. The second-order fluid
4. Quasi-linear models
5. Non-linear models

67
1. Nonlinear Phenomena

A. Normal stress differences in shear flow

Stresses Material functions


vx
" xy
! xy %=
!
" xx # " yy
N1 = ! xx " ! yy $1 =
!2
" yy # " zz
N 2 = ! yy " ! zz $2 =
!2

When a viscoelastic material is subjected to a shear flow, in addition to the viscous


shear stress (xy), there are normal stress differences that occur. The difference
between the xx and the yy component is called the first normal stress difference
and the the difference between the normal stress component on the x-plane and the
normal stress component on the y-plane. The second normal stress difference is
defined in a similar fashion for the normal stress components on the y-plane and z-
plane. Normal stress differences will be one of the first aspect we want to describe.

N1 is typically positive (an exception are liquid crystalline polymers at intermediate to


high shear rates). N2 is of the opposite sign and the ratio -N2/N1 lies in the range 0.01-
0.3. At sufficiently low shear rates, the shear stress will always become proportional to
the shear rate. Similarly the normal stress differences (which are even functions) will
become proportional to the (shear rate)2. The material functions that hence can be
defined are called the primary and secondary normal stress coefficient.

68
A. Normal stress differences in shear flow

The above figure presents the shear and normal stress differences for a solution of
1.2% polyisobutene in decalin (data from Keentok, et al, JNNFM, 1980). N1 is positive
and is the property responsible for rod-climbing. N2 is typically a factor of ten smaller
and has the opposite sign of N1 . The ratio between N1 and N2 is an important check
for the validity of models as it depends very strongly on how nonlinearity is being
described (biaxiality?).

Think about this : why are these non-linear properties?

N1 can be easily measured in rotational devices, e.g. in a cone and plate rheometers
it can be derived from the total axial trust on the cone (see next chapter on
rheometry). The measurement of N2 is much more difficult. It requires measurements
of the pressure gradients along a cone or can be measured using rheo-optical
methods. It is rarely being measured in the industrial laboratories.

69
B. Shear thinning of and 1

The above figure (from Laun, Rheol. Acta, 1978 ) shows the steady state viscosity
and the primary normal stress coefficient for our favorite LDPE melt which weve
already encountered previously. This figure shows that the shear thinning in the
viscosity is typically accompanied by a decrease of the primary and secondary normal
stress coefficients for polymer melts and concentrated polymer solutions. We will
provide a physical explanation for this in the chapter on rheology of polymeric liquids
based on the concept of entanglements.

The data in the above figure were obtained by shifting data to a master curve by using
a shift factor. We will come back to the time-temperature superposition principle later
on.

70
B. Shear thinning shows up in time-dependence

Shear thinning also shows up in time dependent viscosity after starting up the flow at
a constant shear rate. The above figure shows data from Menezes and Greassley
(1980) on a polybutadiene in an organic oil. The lowest shear rate gives us the
behavior that we already can describe in the framework of linear viscoelasticity. As
the shear rate increases, an overshoot appears, at least for deformations that are
sufficiently large and shear rates rates that are sufficiently high. The presence of an
overshoot will be typical for measurements in the shear-thinning regime.

71
C. Relaxation modulus at large strains

Let us see what happens with another LV property, the relaxation modulus, when we
increase the strain amplitude. The above figure gives us the response for the IUPAC
LDPE (data from Laun). At small strains, we find the LVE response which weve
described previously. As the strain is increased, we find a systematic decrease of the
relaxation modulus both the modulus of the shear and the first normal stress
difference.

If you think about this figure it already indicates to us a route to describe non-linear
viscoelastic behavior. The intrinsic shape of the modulus/time curve remains the
same, it is only shifted to smaller values on the vertical axis. Hence a description of
the time effects (the memory function) and the strain effects (a damping function),
would seem to be an attractive route to model NLVE behavior. A (mathematical)
separation of these effects might lead to tractable equations.

72
D. Extensional Thickening in uni-axial flow

1e+7

1e+6
-1
-1
0.1 s
0.5 s
1e+5 -1
1s
(Pa.s)

1e+4
E
!

fiber windup
RME
1e+3 3*!LV

1e+2
0.01 0.1 1 10 100
t (s)

Up to now only nonlinear phenomena in shear flow have been discussed. One of the
annoying things about non-linear behavior is that we cannot necessarily extrapolate
the behavior in shear flow to any other type of deformation.

A number of polymeric fluids exhibits strain hardening in elongational flow. The


above example is for a LDPE (related to the samples studied by Laun in the 70s). At
large deformations, the viscosity deviates from what would be expected on the basis
of linear viscoelasticity. The elongational viscosity drastically increases. This is an
important aspect for processes such as fiber spinning and film blowing.

In the chapter on rheology of polymeric fluids, we will discuss the physical origin of
this strain hardening.

73
E. Phenomena in complex flows

- Contraction Flows : mixed shear / elongational flow


extension

shear

- Hyperbolic flow fields

stagnation point

Up to now only nonlinear phenomena in simple shear or uni-axial elongational flows


have been discussed, but in the introductory lecture we already encountered a wide
variety of phenomena related to non-linear material behavior.

The reliability of a constitutive equation needs to be verified for such complex flows,
as a successful prediction of properties in shear and uni-axial elongation do not imply
that the constitutive model works good.

Hence a comparison of model predictions and stress measurements (by optical


techniques) is crucial for the evaluation of models. Such experiments and calculations
are available in the literature for a number of constitutive equations.

74
2.Tools for Non-Linear Models
Need additional mathematical tools:

A. Deformation measures:

C = FT ! F Cauchy-Green

C "1 = B = F ! F T Finger tensor

B.Time derivative : Upper convected

" $ (!v! )T " A $ A " !v!


!
A=A UC

" = # A + v! " !A
A Material
#t time der.

We already introduced the Finger tensor in the chapter on finite deformation tensors. Physically, the Finger
tensor gives us the relative local change in area within the sample.

#xi #x j
C !ij 1 = "
#x'k #x'k
Another tensor which describes relative deformations is the Cauchy Green tensor. It relates the local
extension ratios with each other.

"x "xk
Both C and C-1 are adequate measure of relative
equations.
Cij = kdeformations
! and are useful to build constitutive
"x' "x' i j

Now we have one more step to take: are the derivatives we know sufficent? Consider the following example
of an expanding duct:

for an external observer the fluid motion is stationary, but for e.g. a polymer molecule in this flow field (an
observer travelling with the stream), it observes a deceleration and hence the time derivative is not zero.

75
2.Tools for Non-Linear Models
B.Time derivative : Upper convected

" $ (!v! )T " A $ A " !v!


!
A=A UC

" = # A + v! " !A
A Material
#t time der.

We know from our courses on Transport Phenomena how to solve this: we use the
Material Time Derivative. This improves the description of what the molecule sees.
It is, however, not sufficient. Remember that solid body rotations should not result in
stresses, neither should the derivatives - neglecting inertia effects (which we can do),
it shouldnt matter how fast a molecule is being rotated.

The upper convected derivative is the derivative in a coordinate system embedded


in the fluid. The coordinate axes are defined such that they deform with material lines
( We can also choose a coordinate axes defined by the material planes of an element
- this would lead to the lower convected derivative). When we use the upper
convective derivative we obtain a description that gives only stresses when a material
element is being deformed. Note that the convective term introduces nonlinear terms!

It is also instructive to note that the UC derivative of the Finger tensor (which
describes relative deformations!) is zero, because of the way that they are both
defined. So we will need UC derivatives of D or to introduce nonlinearity.

76
3. The Second-Order Fluid

$
& = #pI + 2%O D # "1, 0 D+ 4"2, 0 D ! D

Simplest CE capable of predicting a first normal stress difference

Three coefficients needed to characterize material behavior

Only applicable for slow flows

the second order fluid is a logical extenion of the Newtonian CE. We already know
that the term in D.D is only capable of predicting a second normal stress difference.
Hence we need a term giving us a first normal stress difference. We will assume a
small perturbation of the behavior at rest. There is a class of constitutive equations
that we can form in this manner : they are called the retarded motion expansion .
The above equation belongs to this class and is called the second-order fluid. It uses
an additional term containing the upper convective derivative of the rate of
deformation. This term introduces a weak-elastic memory into the equation.

It can be shown that under quite general conditions, that a VE fluid will obey the 2nd
order fluid, provided the flow is sufficiently slow and slowly varying, to ensure that
the deviations from Newtonian behaviour.

Lets first see what this equation predicts in shear flow.

77
2nd order fluid in steady state shear flow
T
)0 0 0& )0 +" 0 & )0 +" 0 & ) 0 0 0 & ) +" 2 0 0&
! 1' $ ' $ 1' $ ' $ ' $
"
D = D * ' +" 0 0 $ " ' +" 0 0 $ * ' +" 0 0 $ " ' +" 0 0 $ = *' 0 0 0$
2' 2' '0
(0 0 0% (0
$ ' 0 0 $% (0 0 0 $% '( 0 0 0 $% ( 0 0 $%

" = # D + v! " !D = 0
D
#t
Homogeneous
Steady state flow

& + xx + xy 0 # &0 )! 0# & )! 2 0 0# & )! 2 0 0#


$ ! 1$ ! $ ! 1$ !
$ + yx + yy 0 ! = ( pI + 2*0 $ )! 0 0 ! + '1, 0 $ 0 0 0 ! + 4'2, 0 $ 0 )! 2 0!
2$ $0 4$
$ 0
% 0 + zz !" %0 0 0 !" % 0 0 !" %0 0 0 !"

$ xy $ xx " $ yy $ yy " $ zz
%0 = !1 = = !1, 0 !2 = = !2, 0
#! #! 2 #! 2

We assume a steady state flow the time derivatives are zero. When we are considering a
homogeneous shear flow, the divergence of the stress tensor will be zero as well (there will be no
gradients in the stress this is the definition of a homogeneous flow).

As you can see, the resulting equation is not all that hard to solve. You can also see how the upper
convected derivative works : when material elements are deformed, a stress is produced. And it produces
a term that gives rise to a first normal stress difference.

Note that we defined the coefficients of the second-order fluid also in a correct manner. In shear flow, the
viscosity is predicted to be constant, and a first and second normal stress difference are predicted, with
constant normal stress coefficients.

Hence the model cannot deal with shear-thinning, but predicts normal stresses. Let us consider the
behaviour in extensional flow.

78
2nd order fluid in steady state uniaxial elongational flow

& + xx 0 0 # & 2)! 0 0 # & 4)! 2 0 0# & 4)! 2 0 0#


$ ! $ ! 1 $ ! $ !
$ 0 + yy 0 ! = ( pI + *0 $ 0 ( )! 0 ! + '1, 0 $ 0 )! 2 0 ! + '2, 0 $ 0 )! 2 0!
2 $ 0
$ 0
% 0 + zz !" $0
% 0 ( )! !" % 0 )! 2 !" $ 0
% 0 )! 2 !"

" E = 3"0 +
3
(
#1 + 2#2 , 0 !
2 ,0
)

2nd order fluid in transient flow?


3
!0 "!
3!0"! +
2
(#1,0 + 2#2,0 )"! 2
xy
E

0 t
0 t

As a first example, solve for a steady state, homogeneous, uniaxial elongational flow. The elongational
viscosity at zero stretching rate equals the Newtonian one, but an extensional hardening is predicted when
-2,0 / 1,0 < 0.5 , where E increases linearly with increasing stretch rate. Note that here we see the power
of the CE, once the three coefficients are known from the behaviour in shear flow we can predict the
behaviour in extensional flow.

As the second-order fluid contains only terms in D and its higher order terms and a UC derivative, it will not
be able to do a good job for predicting time dependent phenomena. For example, for a stress relaxation
experiment the stress will vanish as soon as the flow is stopped. This will be a general result for ANY
retarded motion equation. Above are the predicitions for start-up of shear and elongational flow.

This is what we mean by saying that the second order fluid predicts only weak elastic effcts, some
nonlinear effects are predicted, but time-dependent characteristics, even in the LV regime are not
described. Most flow of polymeric liquids will not be slow enough for the 2nd order fluid or any of its
colleagues to be valid. So well have to abandon this route for writing good CEs.

Our efforts have, however, not been in vain. Any model that describes a smooth evolution of stress and
structure, should, in the limit of slow flows, yield the second order fluid as a limiting case. So apart from LV
behaviour, we have slow flow behaviour as a good test for any non-linear models.

79
4. Quasi-linear Models

A. Differential form: Upper convected Maxwell model

"
! + # ! = 2 $0D

Simplest way to combine time-dependence and elasticity

Non linear because of the UC derivative

but linear in the convected reference frame

Extending the linear viscoelastic Maxwell model into the nonlinear regime can be done by replacing the
material derivative in the Maxwell equation with an upper convective derivative of the stress. This equation
is called the Upper convected maxwell model. The equation is nonlinear as the UC derivative contains the
product of the velocityb gradient and the stress tensor. For dmall strain amplitudes, the nonlinear terms
vanish and a material derivative is all that remains- hence the limit of linear VE is recovered.

Likewise, for very slow flows, the rate is small and Newtonian behaviour is recovered. Analyzing what will
be the first perutbation as thedeformation rates is increased we have to evaluate the upper convected
derivative of the Newtonian equation and :

& $ 2"0 D + second order terms


% %
& $ 2"0 D+ third order terms
Hence, for slow flows, the Maxwell model reduces to a second order fluid with 1,0 = 20 and 2,0=0.
%
& $ 2"0 D # 2"0 ! D+ ...

80
UCM in transient and steady state shear flow

&
( ) # ! " %! # ($v)= 2' D
T
!+% ! + %v # $! " % $v 0
&t

& ! xx ! xy 0 ) &! ! 0 ) & 2 $! ! xy $!


! yy 0) & 0 $! 0)
(! ! + # ( xx xy + ( + = % ( $! 0 0+
0 + " ! ! 0 , " $!
! 0 0
( xy yy + #t ( 0 0 ! + ( 0
xy yy yy
+ 0
( +
' 0 0 ! zz * ' zz * ' 0 0* ' 0 0 0*

Steady state

! xy = "0 #! ! xx = 2 "0 $#! 2 ! yy = ! zz = 0

Again, start from the steady state, homogeneous shear flow. The horrible looking
equation is easily solved. We find a prediction for a constant viscosity, a first normal
stress difference with a constant first normal stress coefficient and a zero second
normal stress difference.

This CE is not yet describing all the phenomena we want, but at least it looks
promising, as we are finding the right limits for VE and slow flows.

81
UCM in transient and steady state shear flow

.
#! xx
! xx + " = 2 "! xy $
#t t=0, no stresses
#! xy .
! xy + " = %0 $
#t
transient flow, predictions of an UCM model
!0=10, " = 1s
25

(
!( t) = !0 1 # e # t / " ) 20

* # $
t
t '- 15

01 ( t) = 2 !0 " ,1 # e " & 1 + ) /


% " ( /. 1
#, !
,+ 10

5
!
#1
0
0 2 4 6 8 10 12 14

t [s]

We are more interested in the transient flows, again writing down the equations in
component form, for a homogeneous flow, you find that the only terms of the stress
tensor differing from zero are given above, and this can be easily solved.

We see that for the shear stress we get the same result as the linear Maxwell model,
BUT! we also predict an evolution of normal stresses in shear flow. The UCM nicely
predicts a gradual increase of the stresses afte inception of flow.

82
UCM in transient elongational flow

2 !0 !0
!E = . + .
1$ 2 " # 1$ " #

10000

1000
e
!

100

10
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

rate (s-1)

Using the same parameters as on the previous page, and using the velocity gradient for uniaxial elongation,
the prediction of the elongational viscosity is obtained under steady state conditions as given above. Note
that this is a quite dramatic non-linear behaviour. The viscosity is predicted to be extremely extensional
thickening, and it rises to infinity when the strech rate approaches 1/2.

In the limit of zero shear rate, the Trouton ratio of three is obtained, followed at slow rates by an evolution
that can be aproximated by a linear dependency on the extensional viscosity on stretch rate, as for the
second-order fluid.

Typically, constitutive equations will differ more amongst themeselves in elongational flow predictions (this
is a strong flow).

Although the UCM model has some of the qualitative features of the rheology of polymeric fluids, it still fails
to describe some essential aspects, such as shear-thiining, N2 and a limited rise of E. The UCM model or
related CEs will do a good job for polymer solutions which are very dilute, such that the polymer molecules
do not overlap. The viscoelastic properties of the solution will be fairly simple and determined by the
properties of the individual molecules. A molecular theory can be constructed that boils down to the UCM
(see part 3).

83
4. Quasi-linear Models

B. Integral form: Lodge model

( )
t
!= $"#
m( t " t' ) C"1 " I dt'

(
# t# t$ )
'0
(C )
t
#1
!= & "e ( # I dt $
2
#% (
Simplest generalization of the LV integral model

the UCM model can be derived from the Lodge model

Results : identical to UCM

The linear Maxwell equation could be written in integral form. The non-linear upper-
convected Maxwell model can also be written in integral form, provided the proper
deformation measure, i.e. the Finger Tensor, is used. By using the fact that the
upper convective derivative of the Finger tensor is zero (the upper convective
derivative is a time derivative in a special coordinate frame whose base coordinate
vectors are stretching and rotating with material lines), the UCM model can be derived
from the Lodge model.

Although equivalent to the differential UCM equation, some problems will be easier to
understand or deal with with integral CE. The fact that real polymeric fluids posses a
distribution of relaxation times is one of them.

The Lodge model, like the UCM, is inadequate in describing most materials because
its elasticity is that of a simple Hookean solid. We will see how we can get a more
general CE.

84
5. Non-linear Models

A. Differential form: of the Maxwell type

1 #
" + " + fC ( " , D) + fD ( " ) = 2G0 D
!

fC ( ! , D) Stress buildup

fD ( ! ) Stress decay

Starting from the UCM, we can try to obtain an equation that modifies the rate at
which the stress tends to build-up and decay. Two functionals will be introduced,
one depending on the stress and rate of deformation tensor - describing the rate at
which stress builds-up and a second one depedning only on the stress tensor that
can accelerate the rate at which stress decays.

Shear-thinning and strain softening can hence be introduced through fc and fd. In
principle, they have a different effect on how non-linearities are introduced, but when
sufficient relaxation times are present similar non-linear effects can be introduced
through either fc or fd.

85
A.1. Differential form: Phan Thien Tanner (PTT)

fC ((, D) = +(D ' ( + ( ' D )


1 &) #
fD (() = exp$ tr ( ! I
* %G "

Linear parameters , G0
Nonlinear parameters ,

Describes : shear thinning


variable N1
Nonlinear time effects
extensional thickening
but : unrealistic oscillations in shear flow
no good description of linear limit

Several forms have been proposed for fd and fc. Each of the equations has
weaknesses and strengths.

One example is the PTT model. The 1 parameter model describes shear thinning,
Non-linear time effects and extensional thickening, without divergence. But some
weird oscillations are obtained during start-up of flow.

The multimode model works much better and can be used to describe a fairly wide
range of polymer melts.

86
A. 2. Giesekus model

fC ( ! , D) = 0
#
fD ( ! ) = ! "!
G0

Linear parameters , G0
Nonlinear parameter

Describes : shear thinning


variable N1, no N2
Nonlinear time effects
extensional not soo good
anisotropic drag!

The Giesekus model is attractive as it contains only one parameter, and still does a
fairly good job. Note that the linear parameters can be obtained from Linear
viscoelastic measurements, the non-linear parameter must be obtained from a non-
linear experiment.

87
B. Really nonlinear integral models

$ [m ( t " t' ,I ]
t
!= 1 C "1
, IIC"1 ) C"1 + m2 ( t " t' , IC"1 , IIC"1 ) C dt'
"#

one can assume m2=0 :


problems N2 = 0, thermodynamics

or assume m2 = m1.

often one can assume factorable mi

Kaye Bernstein-Kearsley-Zappas (KBKZ)

"U "U
m1 = 2 , m2 = 2
"IC!1 "IIC!1

A very general CE for non-linear VE materials which is an extension of the Lodge


model is presented above. Both the Cauchy and Finger tensor are used, this is to
introduce both a first and a second normal stress difference. The simplemodels
assume m2 =0 , there are some fundamental problems.
The experimental data of the stress relaxation data and G(t) in the non-linear regime
show that hey are nearly parallel on a log-log plot, meaning that the effects of strain
and time are factorable.
An expression for the functions m1 and m2 was proposed independently by Kaye and
Bernstein, Kearsly and Zappas in 1962. The functions U are time dependent elastic
kernel functions. Because these functions depend on strain invariants as well as time,
often a lot of experimental data is needed to evaluate these functions, or some
guidance of molecular theories is needed.

Time-Strain separability (factorable mi), simplifies this equation considerably.

88
The above data refers to the LDPE by Laun. The damping function is obtained
experimentally, the approximation with one (dashed line) and two exponentials (sum,
line) is shown. For the single exponential version, only the exponent needs n to be
determined as non-linear parameter. The Wagner model then can be used to predict
the non-linear behaviour, and does a good job for example during start-up of a shear
flow, for both shear and normal stresses. Calculating this evolution requires solution
of an integral equation.

89
B. Really nonlinear integral models

( ) (
m1 t ! t' , IC!1 , IIC!1 = m( t ! t' )h IC!1 , IIC!1 )
h is called the damping function

Wagner model

( ) (
h = f1 exp !n1 Ic !1 ! 3 ! f2 exp !n2 Ic !1 ! 3 )
h( " ) = exp( !n" )

When the strain-time separability can be applied, the only nonlinear material
function needed to predict the behaviour is the damping function. It will, in general
depend on the invariants of the Finger tensor. In some case, a simple exponential or
for higher strains a sum of two exponentials is sufficient. This model was derived by
Manfred Wagner (1976).

90
Conclusions

yield
Gen. Newt

NLVE
N 2nd E
0 OF
Integral
Differential
..

LVE
#
De De =
"! !1

91
Conclusions of part 1:
Material functions that we can use to describe viscoelastic behavior and Non-
Newtonian behavior have been rigorously introduced.

Bovenstaande grafiek - uit het boek van Morrison -

92
Outline

1. Introduction

2. Shear rheometry : drag flows

3. Rotational rheometers : practical design

4. Shear rheometry : pressure driven flows

5. Capillary rheometers : practical design

6. Extensional rheometry

(7. On-line rheometry)

93
Introduction

Why ? 1. Input for CE : material functions


2. Quality control
3. Simulate industrial flows

A rheometer is an instrument that measures both stress


and deformation.
indexer
viscometer

What do we want to measure?


- small strain (LVE) functions
- large strain deformations
- steady state data

what are the difficulties? Many!

94
Introduction: classification

Kinematics : shear vs elongation

homogeneous vs non-homogeneous vs complex flow fields

type of straining:

- small : G(), G(), +(t), -(t), G(t), y


- large : G(,), G(,), +(t,), -(t ,), G(t ,), (t,)
- steady : ( !), 1( ! )

shear rheometry : Drag or pressure driven flows.

95
Shear flow geometries

Drag flows Pressure driven flows

Sliding plates

Capillary
Couette

Slit
Cone and Plate

Annulus
Parallel plates

96
Requirements for shear rheometers:

1. Flow can be analyzed


2. Laminar flow
3. No secondary flow
4. Possibility to vary and shear rate
5. Boundary effects : negligible or correctable
6. Temperature homogeneity and control
7. Large strain expts: - homogenous
- no inertia
- fast response

97
1. Drag flows : Sliding plate rheometer

Fy v0
!=
h
Fx
Fx
h " 12 =
LW

L " 22 =?

- difficult to keep plates parallel


- difficult to achieve steady shear
- edge effects very important, local measurements not easy
- sample must be viscous or flows out
- buckling occurs when sample too viscous
- not very practical

98
Drag flows : Couette geometry

M. Couette (1890)

99
Drag flows : Couette geometry

2R0 Concentric cylinders, inner one rotating :


2Ri
1. Steady, laminar, isothermal flow
2. Negligible gravity and end effects
3. Symmetric in
4. vr=vz=0 and v=r

L Equations of motion:
1 !(r" rr ) " $$ v2
r: % = %# $
r !r r r

$:
(
! r 2 " r$ )= 0
!r
% !p
z: + #g = 0
!z

100
Drag flows : Couette geometry

Boundary conditions 2R0

1. v=Ri at Ri 2Ri

2. v=0 at Ro

Shear stress
L

!:
( 2
" r # r! )= 0 % # r!
C
= 2)
'

"r r )) M
( # r! =
) 2$ & L & r 2
M )
# r! (Ro) & (2$ & Ro & L) = )*
Ro

Normal stress :
- causes rod-climbing
- stress difference @ Ri and Ro : hard to measure

101
Drag flows : Couette geometry

-verry narrow gaps: shear stress is approximately constant.


shear rate is approximately constant

" #R Ri
!= if > 0.99
Ro $ Ri Ro

Ro + Ri
R =
2

-if Ri/Ro <1: shear stress is not constant over the gap.
shear rate distribution

"#
!= r
"r
d#
! = 2$
d$
# (Ri ) $(Ri ) !
#(Ri ) = % d# = % d$
0 $(Ro ) 2$

d#
2$(Ri ) = ! ($(Ri )) & ! ($(Ro))
d$(Ri )

102
Drag flows : Couette geometry

+ - if Ri~Ro constant shear rate


- large measurement surface, good accuracy
- quite good for settling materials

- - difficult to fill for high viscosity materials (bubbles)


- secondary flow at high - limits max. shear rate
- heavy geometry : inertia!
- no good normal stress measurement possible
- thermal control / viscous heating
- solvent evaporation
- end effects need to be reduced
or corrected for:

conicylinder, recessed bottom or Double Couette

103
Drag flows : Cone and plate

Probably most popular geometry Mooney (1934)

1. Steady, laminar, isothermal flow


2. Negligible gravity and end effects
3. Spherical boundary liquid
4. vr=vz=0 and v(r,)
5. Angle < 0.1 radians

Equations of motion:

r:
( 2
1 ! r " rr )
&
" $$ + " %%
= & #
v $2
r2 !r r r

1 !(" r$ sin $) cot $


$: & ' " $$ = 0
r sin $ !r r
1 !" $% 2
%: + cot $ ' " $% = 0
r !$ r

104
Drag flows : Cone and plate geometry

Boundary conditions
1. v(/2) = 0
2. v(/2-)=rsin(/2-)r

Shear stress

!:
( )
1 d " #! 2
+ cot ! & " #! = 0 ' " #! =
C *
( Cte ,
r d# r 2
sin # ,
, 3M
+ " #% =
, 2$ & r 3
2$ R
M = ) ) r 2 " #! drd! ,
0 0 ,-

Independent of fluid characteristics because of small angle!

105
Drag flows : Cone and plate geometry

Shear rate: is also constant throughout the sample: homogeneous!

vr " $r " "


!= = = %
h(r ) r $ tg (# ) tg (# ) #

Normal stress differences: total trust on the plate is measured Fz

"R 2
Fz =
2
(
# $$ ! # %% )
2Fz
N1 =
"R 2

106
Drag flows : Cone and plate geometry

+ - constant shear rate, constant shear stress -homogeneous!


- most useful properties can be measured
- both for high and low viscosity fluids
- small sample
- easy to fill and clean

- - High visc: shear fracture - limits max. shear rate


- low visc : centrifugal effects/inertia - limits max. shear rate
- settling can be a problem
- solvent evaporation
- stiff transducer for normal stress measurements
- shear and viscous heating

107
Drag flows : Parallel plates

Again proposed by Mooney (1934)

1. Steady, laminar, isothermal flow


2. Negligible gravity and end effects
3. cylindrical edge
4. vr=vz=0 and v(r,z)

Equations of motion:

1 !(r" rr ) " $$ v2
r: % = %# $
r !r r r
!(" $z )
$: =0
!z
!(" zz )
z: =0
!z

108
Drag flows : Parallel plates

Shear rate: is not constant throughout the sample

vr " #r
!= =
h h

Shear stress M $ d ln M '


!= &1 + )
2"R 3 % d ln #! R (

Normal stresses
Fz " d ln Fz %
N1 ! N2 = 2 $
2+ '
(R # d ln )! R &

109
Drag flows : Parallel plate geometry

+ - preferred geometry for viscous melts - small strain functions


- sample preparation is much simpler
- Shear rate and strain can be changed also by chaging h
- determination of wall slip easy
- N2 when N1 is known
- edge fracture can be delayed

- - non-homogeneous flow field (correctable)


- inertia/secondary flow - limits max. shear rate
- edge fracture still limits use
- settling can be a problem
- solvent evaporation
- shear and viscous heating

110
Rotational rheometers : practical designs

Stress controlled strain controlled

Transducer

Normal force control


Drag cup motor

Torque control

Position sensor(s) Platform


optical Air bearing

Servo Controlled
Geometries Motor
environments

111
Pressure driven flows : capillary rheometry

Hagen (1839)
Poisseuille (1840)

Non-homogeneous flow
typically for high shear rates

simpler to operate
yet yield precise data!

112
Pressure driven flows : capillary rheometry

Q
R 1. Steady, laminar, isothermal flow
2. No slip at the wall, vx =0 at R=0
3. vr=v=0
4. Fluid is incompressible, f(p)

! r
Equation of motion:
Z

1 !(r" rz ) !p
L z: # =0
r !r !z

Only a function of r
Only a function of z

1 d (r! rz ) dp
=
r dr dz

113
1 d (r! rz ) Po " PL
=
r dr L
Po " PL r C2
! rz = # +
L 2 r
C2 = 0 because ! rz $ % @ r =0

Po " PL R
! rz (R ) = ! w = #
L 2

r
! rz (r ) = ! w
R Note : independent of fluid properties

114
R
Q = 2) ! v z (r )rdr Shear rate calculation from Q
0

R
R dv z
Q = 2 )v z r 0
" 2) ! r 2 dr Integration by parts + assume :no slip
dr
0

R R substitute r and dr
r= * and dr = d*
*w *w

*w # 2
R
dv z R & dv z R
Q = "2 ) ! r 2 dr = "2 ) ! % *( d* Typical trick to deal with inhomogeneous flows:
0
dr
0 $ * w '
dr *w changing of variables

Q* w 3 *w 3Q! w 2 !w 3 dQ
= " ! (* ) = % ! w 2 dv z
2 dv z
d* + $
)R 3
0
dr "R 3
"R 3 d! w dr !w

Differentiate with respect to w


3Q 3Q d ln Q
using Leibnitzs rule #! w = % dv z = + $
dr !w 4 "R 3
4 "R 3 d ln ! w

#! a & d ln Q )
#! w = (3 + +
4 ' d ln ! w *

115
!a # d ln Q &
Weissenberg-Rabinowitsch Correction !w = %3 + (
4 $ d ln " w '

Correction factor accounts for material behaviour

Physical meaning: 2

n=1
n=0.7
n=0.3

e.g. power law fluid n=0.01

>
!a "

z
1% 1

!w = $3 + '

/ v<v
z
4 # n&
Steepness of the velocity profile changes
Shear rate is increased with respect to 0
0 1
the Newtonian case: r/R

4Q
!a =
"R 3

116
Rabinowitsch Weissenberg correction : be careful - requires numerical differentiation

1000

Raw data
"corrected data"

100
[Pa.s]
a
!, !

10
typical data of a PP sample @ 230C

10 100 1000 10000

shear rate [s-1]

Schmmers method : !( "! ) = !a ( "! a ) "! = 0.83 "!


a

117
Are these conditions met ? 1. Steady, laminar, isothermal flow
2. No slip at the wall, vx =0 at R=0
3. vr=v=0
4. Fluid is incompressible, f(p)
Problem 1: Melt distortion

Melt fracture typically occurs at w 105 Pa

118
Are these conditions met ? 1. Steady, laminar, isothermal flow
2. No slip at the wall, vx =0 at R=0
3. vr=v=0
4. Fluid is incompressible, f(p)
Problem 2: Viscous heating

.
!"R2 #
Na = 4k

119
Are these conditions met ? 1. Steady, laminar, isothermal flow
2. No slip at the wall, vx =0 at R=0
3. vr=v=0
4. Fluid is incompressible, f(p)
Problem 3: Wall slip

4vs
"! a = "! + @ ! = cst
R

120
Extensional Rheometry
Extensional viscosities can provide an
indication of chain structure.

thickening: flexible, dilute chains

3
thinning: rigid systems

thickening: branched melts

3
thinning: linear melts

Why is extensional rheometry important?

Most processing flows are complex superpositions of flow types and many contain substantial
amounts of extensional character.

The extensional viscosity is qualitatively different than the shear viscosity and complete
characterization should include its measurement.

Extensional flows are extremely difficult to realize in the laboratory. The available techniques
can be divided into methods for melts and methods for solutions.

121
Extensional Viscosity of Melts
Extensional flows for melts can be generated more easily than
solutions because these systems are sufficiently viscous to
allow them to be mechanically stretched. Below is the Meisnner
device (Rheol. Acta 33, 1-21, 1994), other examples include the Mundstedt
device.

sample

frit
air support

The major advantadges of the Meissner device are that effects of gravitation are
nicely balanced by the air flow and that the deformation field is homogeneous.

Nevertheless, sample preparation is not at all easy, slippage at the belts can occur
and the range of deformation rates and forces is typically limited.

122
Macoskos Wind-up Technique
J. Rheol. 40, 473481 (1996)

!R D
"! = ,R D windup drum radius,
L0
L 0 distance between clamp and drum

M exp("! t )
#11 $ # 22 = ,M torque,
!R D R 02
R 0 initial radius of the sample

This technique goes back to the 60s but was reintroduced and improved by Macosko.

The advantages of the technique are that it can be mounted on a rotational


rheometer, high extension rates can be achieved and the sample preparation is easy.

The deformation field is however not perfectly homogenous (one end is clamped in -
think about this) and sagging under the effect of gravitation cannot be avoided.

123
LDPE (ATOFINA) @ 190C

1e+7
0.1 s-1

0.5 s-1
1 s-1

1e+6

3*!
[Pa.s]
E
!

1e+5
3*!afschuiving

1e+4
0.1 1 10 100

t [s]

These are typical data taken on a commercial LDPE sample in our lab using the FWU
technique. The pronounced strain hardening that can be observed makes this a
suitable material for e.g. film blowing.

124
Capillary Breakup methods

125
Current Research Topics

- Linking rheology and structure:


Rheo-optics, Rheo-scattering,...

- Rheology of complex fluid interfaces

- Microrheological methods

- Rheology of soft matter composites

- Role of rheology in biological systems

126
Current Research Topics

- Rheology of soft matter composites

127
- Linking rheology and structure:
Rheo-optics, Rheo-scattering,...

128
- Rheology of complex fluid interfaces

129
Aim of this lecture:

1. Qualitative insight of the behavior of non-Newtonian fluids.

2. Discuss measurement techniques for a given problem

3. Demonstrate link rheological properties // microstructure

130

Das könnte Ihnen auch gefallen