Viscous Dissipation

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 63

Viscous Dissipation in Shear Flows of

Molten Polymers

HORST HENNING WINTER

Institut f u r Kunststoff(echno1ogie. Universitat Stuttgarr, Stuttgart, West Germany

I. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
A. System of Equations. . . . . . . . . . . . . . . . . . . . . . . . . 207
B. Thermal Properties . . . . . . . . . . . . . . . . . . . . . . . . . 209
C. Rheological Constitutive Equation. . . . . . . . . . . . . . . . . . . 211
11. Shear Flow (Viscometric Flow). . . . . . . . . . . . . . . . . . . . . . 212
A. Thermal Boundary Condition. . . . . . . . . . . . . . . . . . . . . 222
B. Steady Shear Flow with Open Stream Lines. . . . . . . . . . . . . . . 221
C. Shear Flow with Closed Stream Lines . . . . . . . . . . . . . . . . . 250
111. Elongational Flow; Shear Flow and Elongational Flow Superimposed
(Nonviscometric Flow) . . . . . . . . . . . . . . . . . . . . . . . . . 260
IV. Summary, . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
Nomenclature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264

I. Introduction

Polymer processing and applied polymer rheology occur at relatively high


temperatures and often at high temperature gradients. In molten polymers,
large stresses are required to maintain the flow and, additionally to convective
and conductive heat transfer, temperatures essentially depend on viscous
dissipation, i.e., on conversion of mechanical energy into heat. Velocity and
temperature fields influence each other: the temperatures influence the flow
through the temperature-dependent rheological properties, and the velocities
influence the temperatures through convection, through dissipation, and
through anisotropical effects (which are investigated very little) on the
thermal properties.
205
206 HORSTH. WINTER

Research in rheology and in thermodynamics related to heat transfer


problems is mostly done separately, by rheologists on molten polymers or
polymer solutions at constant temperature, and by thermodynamicists on
polymers at rest. The difficult task of combining the two areas is left to the
polymer engineers (see for instance [1-41). A number of assumptions have
to be introduced into the heat transfer analysis before applied problems can
be solved.
The different flow problems involving heat transfer and viscous dissipa-
tion can be classified in groups as shown in Table I. Each of the groups is
characterized by different rheological phenomena, and one has to choose
very different rheological constitutive equations to describe them. The two
main groups are channeljow (includingflow geometries with partly solid and
partly free boundary) and free surface flow (with no solid boundary). In
polymer processing, channel flow of molten polymers occurs in a large
variety of flow geometries. The polymer is forced through a channel by a

-
pressure gradient (flow in an extruder die, for instance),or it is dragged along
by a moving wall (rotating screw in a stationary cylinder, for instance). Very
often both types of flow are superimposed on each other. Free surface flow
for example occurs in film blowing or fiber spinning.
TABLE I
CLASSIFICATION OF HEATTRANSFER
AND VISCOUS DISSIPATION
IN MOLTENPOLYMERS

heat transfer and viscous


dissipation in molten polymers
hfree surface flow
channel flow
(shear free flow)
shear flow shear flow and
elongational flow
superimposed
shear flow with shear flow with (non-viscometricflow)
open streamlines closed streamlines

For rheological reasons, channel flow problems are subdivided into shear
jow (also called viscometric flow) and nonuiscometric flow. The separation
of the shear flow problems into one group with open stream lines and one
with closed stream lines has to be made since their thermal development is
different.
Throughout the first section, the heat transfer problem will be considered
in general, i.e., the relevant equations are listed in a general form and the
properties are described. The rheological properties of the molten polymers
have to be formulated differently according to the various flow types since
IN FLOWING
VISCOUSDISSIPATION MOLTENPOLYMERS 207

the length of the following sections is supposed to reflect the degree of under-
standing of the respective flow and heat transfer problems.
In Section 11, heat transfer in shear flow will be analyzed. A large emphasis
will be laid on replacing the commonly used idealized boundary conditions,
i.e., constant wall temperature or constant wall heat flux (with the limiting
case of the adiabatic wall), by more general conditions. In practical applica-
tions, the idealized conditions will rarely occur; actually it is difficult to
achieve them even in especially designed model experiments. To make the
analysis applicable, heat transfer in a flowing polymer should not be studied
separately inside the fluid, but together with the surrounding wall.
In this analysis the heat transfer at the wall is described by an outer tem-
perature difference (temperature of the surroundings minus temperature at
the boundary) and the Biot number, which otherwise has been used success-
fully for describing the boundary conditions for temperature calculations in
solids. The Biot number is appropriate for describing boundary conditions
between isothermal and adiabatical, as they occur in real processes. Addi-
tionally, the thermal capacity of the walls is included in the analysis by
introducing the capacitance parameter C .
Heat transfer in viscometric flow has been studied quite extensively in the
literature, and at the present state it seems to be necessary to show the many
common aspects of the different studies. Thus, as the main goal of this study
a unifying concept will be developed. This concept makes it possible to
comprise the most important shearjow cases into a single one, which can be
solved with one numerical program.
For Section I11 on nonviscometric flow in channels and flow with free
boundaries, the description will not go much further than stating the prob-
lem, showing the present methods of solution, and listing references. Since
nearly all of the results in this report are on shear flow, the title is taken to
be “in shear flows” even if the problem is stated in a general form and Section
I11 is on nonviscometric flows.
Heat transfer in non-Newtonian fluids at negligible viscous dissipation is
not included in this report (see instead [5-7]), although it can be treated as a
limiting case of the corresponding flow with viscous dissipation.

OF EQUATIONS
A. SYSTEM
The problems are governed by the equations describing the conservation
of mass
appt +
v * (pv) = 0 (1.1)
and the conservation of energy
p DelDt = V ( k V T )
9 + o :Vv, ( 1.2)
208 HORSTH . WINTER

by the stress equation of motion


p DvlDt = V * u + pg, (1.3)
and by the constitutive equation which will be described below, together
with the appropriate flow geometries. The three equations above are derived
and tabulated in textbooks (see for instance [S]) for different coordinate
systems.
a/a denotes the partial and DID the substantial derivative; V is the “nabla”
operator. Density p and thermal conductivity k are properties of the fluid.
Velocity v, internal energy e, temperature T , time t, and stress CT are the
variables. The stress CT is defined in such a way that the force on the positive
side of a surface element of unit area and normal vector n is n 0.
The equation of energy says that the rate of gain of internal energy per
unit volume ( p De/Dt) is equal to the rate of internal energy input by conduc-
tion per unit volume V (k V T ) plus the rate of work by the stress on the
volume element u :Vv, which is being partly stored and partly dissipated
during the flow. For heat transfer studies, the internal energy has to be
defined in terms of the fluid temperature and the strain and stress variables.
IncornpressibleJiuid: In rheology the fluid is usually supposed to be incom-
pressible (even when properties such as the viscosity are allowed to depend
on pressure). The flow geometry, the temperature, and the rheological prop-
erties of the fluid determine the stress completely, except for an arbitrary
added isotropic pressure [9]. Therefore, the stress is commonly separated into
an arbitrary pressure p and the extra stress 7 , which is defined in the rheolog-
ical constitutive equation, viz.
a = -pd+z. (1.4)
6 denotes the unit tensor.
In some flow problems, it is convenient to define the isotropic pressure p
to be equal to one of the normal stress components in a certain coordinate
system ( p = -ell, p = -crz2, or p = - L T ~ ~ )while
, in other flow problems
it might be preferable to define p = -(trace a)/3.
For an “incompressible”fluid, the change in internal energy and the work
of the stress per unit time are determined by
p DelDt = cp DTIDt,
u:Vv = - p v * v + 7:vv, (1.5)

and Eq. (1.2) becomes


cp DTIDt = V * ( k V T ) + z:Vv. (14
The specific heat capacity c is defined as the thermal energy needed per
unit mass and Kelvin degree for changing the temperature of a material.
Since the density is taken to be constant, c has to be measured at constant
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 209

density. If the fluid were really incompressible, the specific heat should be
the same for measurements at constant density (c,) or at constant pressure
(c,). From thermodynamic data at rest (Eq. (l.ll)),however, one finds that
c, and c, of polymer melts differ by about 10%.
Compressible fluid: There are difficulties in relating strain and stress in
deforming materials that are slightly compressible. One commonly assumes
that the deformation can be separated into two parts: a deformation at
constant density and the volume change [lo]. Neglecting the influence on
each other, the deformation at constant density is described by the consti-
tutive equation, and the density of the flowing polymer is determined from
equilibrium data p(T, p ) measured on the fluid at rest (taking p = -(trace
4/31.
There also seem to be difficulties in defining the internal energy e of a
compressible flowing fluid: one assumes that e can be described in terms of
p and p only, independently of the other stress and strain variables (see for
instance [81):
e = e(p, P).
Applying this relation to the flowing polymer melt, the substantial derivative
of the internal energy then becomes
De
PK= - P V *V + € TDP
-+
Dt
DT
pep-
Dt
(1.7)

with E = -p-l(i?p/aT),, the coefficient of thermal expansion. E and c, are


evaluated from rest data at temperature T and the “pressure” p = -(trace
4 3 . The equation of energy takes the form usually shown in the literature
[ll]:
PC - = V .(k V T ) + ET- + t:Vv.
DT DP
Dt Dt

From this equation, calculated temperature fields in channel flow (see


Fig. 13, p. 244) show large temperature decreases due to cooling by expan-
sion. The assumptions made above (concerning the density changes and the
internal energy) are rather severe, and further experimental studies are
needed to investigate their validity.

PROPERTIES
B. THERMAL
The properties in the analysis are the density p, the specific heat c, and the
thermal conductivity k : the thermal diffusivity is defined as a = k/pc. Rheo-
logical properties are defined separately in the constitutive equation.
In a stationary fluid, the density p ( p , T ) is a function of pressure and
temperature. It can be described by the equation of Spencer and Gilmore
210 HORSTH. WINTER
TABLE I1
CONSTANn OF EQ. (1.9) MEASURED
MATERIAL BY
SPENCER
AND GILMORE [121

polymer b* m3/kg] p* [N/m'] W kg/g-mole]

Id.PE 0.875 3.275 x 10' 28


PS 0.822 1.863 x 10' 104
PMMA 0.734 2.157 x 10' 100
CAB 0.688 2.844 x 10' 54

[12]:
( l / P - b*)(P + P*) = RT/W, (1.9)
where b*, p*, and W are material constants. Their values are tabulated
(Table 11) for some examples of the most widely used polymers; R = 8.314
[J/K g-mole] is the gas law constant.
From Eq. (1.9) one can evaluate the term ET of the equation of energy
ET = 1 - pb*; (1.10)
since b* is always smaller than p-l, the dimensionless product CT adopts
positive values smaller than unity.
The density generally is measured on the polymer at rest and in thermody-
namic equilibrium. Dynamic measurements by Matsuoka and Maxwell [131,
however, show a very delayed response of polyolefines to sudden pressure
changes. Thus, the use of equilibrium density data restricts the analysis to
flows of slowly changing pressures. The reaction to temperature changes is
similarly delayed [141. Additionally the flow might influence the density.
The specific heat commonly is measured at constant pressure. Using the
equation proposed by Spencer and Gilmore, Eq. (1.9),one can determine the
c, from specific heat data at constant pressure
specificheat at constant density.~

C, = cP - R/W. (1.11)
The thermal conductivity k and the specific heat capacity cp are slowly
varying functions with temperature and they also depend on pressure. In
flowing polymers the thermal conductivity possibly varies with direction. For
most polymers the temperature dependence can be expressed in a linear form
k = E(1 + uJT - To)),
(1.12)
cP = Tp(l + a,( T - To)).
-
k and Fp are values at some reference condition (temperature T o ) ,while ak
and a, are the temperature coefficients, which might be positive or negative
IN FLOWING
VISCOUS DISSIPATION MOLTEN
POLYMERS 21 1

TABLE 111
THERMALCONDUCTIVITYAND SPECIFIC HEAT
CAPACITY
OF SOMEMOLTEN POLYMERS AT
TEMPERATURESTo (see [ 15 -231)

To (‘P k
Polymer rC] [lo3 Nm/kg K] [N/K s]

1.d.PE 150 2.51 0.241


h.d.PE 150 2.65 0.255
PP 180 2.80 -

PVC 100 1.53 0.166


PS 150 2.04 0.167
PMMA 150 - 0.195

depending on the polymer in question and on the temperature range.


Table I11 shows some values of k and c p .More detailed data on the properties
can be found in references [15-231, for instance.

C. RHEOLOGICAL
CONSTITUTIVE
EQUATION
For a large number of fluids, which can be regarded as incompressible,
the stress can be described by the Stokes equation
u = -p6 + qy, (1.13)
the simplest tensor generalization of Newton’s law of viscosity. Here p is
the isotropic pressure, and i, = (Vv) + (Vv)’ is the rate of strain tensor; the
viscosity q depends on temperature and on pressure, but not on the time t or
on any kinematic quantities such as 9. Fluids that show this behavior are
called Newtonian.
The Stokes equation has been generalized by taking the viscosity q($) to
be a function of the second invariant of the rate of strain tensor [24] :
u = -p 6 + q($)i,; i, = (+i,:i,)l’Z. (1.14)
This equation defines the “generalized Newtonian fluid.” It has been applied
quite successfully to molten polymers in steady shearjlow for calculating the
shear component of the stress tensor in an appropriate coordinate system;
however, the normal stress components calculated from this equation are
known to be unrealistic for molten polymers. In general, the equation might
be misleading in its tensor form because it does not allow one to calculate
meaningful stress components in arbitrary coordinate systems. The appro-
priate statement of the rheological equation for molten polymers in steady
shear flow is given by Criminale et al. [25]; their equation will be applied in
Section 11.
21 2 HORSTH.WINTER

The rheological properties of elastic fluids (such as molten polymers) at


any given position depend on the strain and temperature history of the fluid
elements when they arrive at that position, independently of the history of
neighboring particles. Translational and rotational movements do not in-
fluence the stress [9].
Depending on the type of flow, the rheological behavior of molten polymers
is more or less different from the behavior of Newtonian fluids. Up to now
there exists no general constitutive equation to describe all the phenomena
known for a given polymer melt. Additionally, the temperature effects on the
rheological properties have not been studied at all or only at different levels
of homogeneous temperatures. The constitutive equations used will be stated
in the beginning of Sections I1 and 111.

11. Shear Flow (Viscometric Flow)

In shear flow at constant density, material surfaces move “rigidly” (i.e.,


without stretching) across each other. These surfaces are called shear suduces
[26]. Pipe flow, which is an example of shear flow, sometimes is called tele-
scopic flow since its rigidly moving surfaces are concentric cylinders. In
Fig. 1 the deformation of particle Po at the origin is described by the relative
motion ofneighboring planes. On the left (Fig. la) an orthonormal coordinate
system is chosen, such that
x1 = direction of shear,
x2 = direction of velocity gradient,
xg = neutral direction.

shear direction D x21 ,shear direction

tan E = dtan E,J*+ [tan c3d2 ‘

FIG.1 . Unidirectional shear flow in Cartesian coordinates. Shear flow in the 1-3 plane as
superposition of shear flow along x, and along xj.
DISSIPATION
VISCOUS IN FLOWING
MOLTENPOLYMERS 213

The direction ofshear (or shear direction) and related quantities are defined
following [26]: an orthonormal coordinate system is chosen to have its x1
axis and its x3 axis in a shear surface (see Fig. 1). The x2 axis is perpendicular
to the shear surface; it projects point Po onto neighboring shear surfaces.
During shear flow, shear surfaces move past each other, and the normal
projection of Po draws a line on neighboring shear surfaces. This line is
called the shear line. (Note: In the examples of Fig. 1, the shear surfaces are
planes, and the shear lines are straight.) The tangent to the shear line at time t
is defined to be the shear direction at time t, and the angle of the shear line
with the x1 coordinate is the shear angle cp; if the direction of shear remains
constant with time, the flow is called unidirectional. For most applications,
the direction of shear is identical with the direction of flow; an exception is
the othogonal rheometer [26, p. 761, for instance.
The shear rate, the extra stress, and the temperature are supposed to be
uniform in the direction of shear, but they may change perpendicularly to
the direction of shear (even within shear surfaces). If one allows for shear in
the 3 direction (Fig. lb) additionally to the shear in the 1 direction, the shear
direction is in the 1-3 plane (Fig. lc); for some steady shear flow geometries
(as in helical flow), it is convenient to choose a global coordinate system with
the velocity vector in the 1-3 plane and the velocity gradient normal to the 1-3
plane. The matrix of the rate of strain tensor becomes

[tl = [
Pl2 2:

Q23
:23]? (2.1)

and the second invariant of the rate of strain tensor defines the shear rate:

In unidirectional shear flow at constant temperature T o ,constant pressure


p,,, and constant volume, the stress is given by the shear rate Q(t) and the
three shear-rate-dependent viscometric functions [25,26] :
viscosity q ( j t T o ,po),
first normal stress coefficient m , T o ,po),
second normal stress coefficient $2(QCm, T o ,po).
Here symbolizes the “history of shear rates” to which the medium was
submitted up to the present time t.
Up to now, there do not seem to exist heat transfer studies on shear flow
in general; however for steady shear flow, the publications are numerous. In
Section II.C.3 an example of heat transfer in unsteady unidirectional shear
flow will be shown.
214 HORSTH. WINTER

Steady Shear Flow and Shear Viscosity


For unsteady shear flow the shear direction and/or the value of the shear
rate change with time; the shear surfaces, however, are maintained. If the
shear direction and the shear rate are kept constant over some time (cp =
const; j~ = const), the shear stress approaches a constant value, i.e., the
viscosity adopts a constant value:
?(t,To,P o ) = lim ?(?‘-m,
t+m
To,P o ) , (2.3)

called the viscosity, the shear viscosity, or the apparent viscosity. The flow
becomes steady (unidirectional)shearpow. The index on To and po indicates
that the viscosity is dejned at constant temperature and pressure.
The elastic properties of the fluid are represented in steady shear flow by
the two viscometric functions t,bl, t,b2. These functions, however, do not have
any influence on heat transfer and viscous dissipation; this will be shown in
the following.
The only two terms in the system ofequations (Eqs.(1.1)-(1.3)) that contain
the stress are a:Vv and V u. The rate of work by the stress a:Vv, which in
steady shear pow is dissipated completely, sometimes is called the dissipation
function. Taking the coordinate system of the shear flow, one can evaluate
the two terms [8, p. 7381; u:Vv is a scalar
0:Vv = 212?12 + 2 2 3 7 2 3 , (2.4)
-
and V u is a vector with the three components

[V.U]l =

[V-a], =

[V .I3 =

The stress u has been decomposed into p and z as shown in Eq. (1.4).If x1
-
is the shear direction (Fig. la), the 1component of V u is used for calculating
the velocity and the pressure gradient; due to the symmetry of the flow
(at/ax, = 0); and since 7 1 3 = 0, the 1 component reduces to

[v*u]l = --a P
ax,
+ -.ax,
a212

For a shear direction in the 1-3 plane (Fig. lc),the 1 and the 3 components are
used for calculating the velocity and the pressure gradient; since az/ax, = 0
IN FLOWING
VISCOUSDISSIPATION MOLTENPOLYMERS 215

and &/ax3 = 0, they reduce to

[V.UIl = --aaPx , + -,a x ,


az12
[V*6I3 = --
ax3
+-
ax,
(2.7)

For calculating the velocity, the pressure gradient, and the rate of work by
the stress, one finds from Eqs. (2.4), (2.6), (2.7) that one needs only the shear
components z12,zZ3of the stress matrix
212 = q(?, 912, 723 = q(?, j23* (2.8)
The shear component 213 = sin 2q1($~+ $,)p2/2 and the normal stress
components do not contribute to the analysis. The viscosity q is the only
rheological property needed for solving heat transfer problems in unidirec-
tional steady shear flow. Thus the heat transfer analysis of Section I1 is not
restricted to purely uiscousftuids, even if the normal stresses are not mentioned
further.
The pressure and the normal stresses can separately be determined from
the pressure gradient, the 2 component of the stress equation of motion
(which contains [V * el2),and the appropriate boundary conditions.
Some typical curves of viscosities referring to different temperatures are
shown in Fig. 2 [27]. At low shear rates (p < 10 s-’) one measures the vis-
cosity in Couette or in cone and plate rheometers, and at high shear rates
( j ~> 1 s-’) one uses a capillary or slit viscometer; the temperature in the
test section is kept as uniform as possible.

10’ 10’ 103 l@


shear rate 7 Is-’]
FIG.2. Typical viscosity curve of molten polymer (low density polyethylene) measured by
Meissner [27].
216 HORSTH. WINTER

The pressure and temperature dependence of the viscosity usually is


described by the corresponding coefficients

pressure coefficient

temperature coefficient = -1 (3)


. (2.10)
v aT b,P

(Note: Some authors define the temperature coefficient p at constant shear


stress instead of at constant shear rate.) In the expression for the viscosity,
the variables are separated. The pressure dependence is described by an
exponential function. For incorporating the temperature dependence one
may take just an exponential function
v(P, T, PI = f ( P 9 P ) exp( - BT) (2.1 1)
or use an Arrhenius type expression

V(P, T, P) = f(P, P ) e x p ( E / W (2.12)


whose “activation energy” E has been reported to be a material constant
over wide temperature ranges [28, 291. The temperature coefficient of the
viscosity at temperatures around To can then be determined as (if one
expands (EIRT) around To)
P(T0) = E/RTo2 (2.13)
i.e., for constant E, the temperature coefficient is proportional T-’. The
temperature dependence of P normally is neglected in analytical studies,
which is acceptable if the deviations AT from the temperature level T o are
not too large. The relative difference of the two expressions, Eqs. (2.11) and
(2.12),is

= 1 - exp[-fl(AT)2/(To + AT)]; (2.14)


as an example AT = 10 K, fl = K-’, T o = 400 K gives a relative
difference of less than 0.25%.
At very low shear rates (p < 10-1-10-3 depending on temperature,
polymer, molecular weight distribution) the viscosity adopts a value inde-
DISSIPATION
VISCOUS MOLTENPOLYMERS
IN FLOWING 217

pendent of shear rate, the zero viscosity qo(T,p ) (see Fig. 2). At medium and
high shear rates (jJ > lo), as they occur in polymer processing, the viscosity
curve q ( j ) is nearly a straight line in the log-log plot. Ranges of the curve can
be approximated by a power law [30] which will be formulated as

(2.15)

The power law exponent is different for different ranges of i), T , p ; for molten
polymers, the value of m is between 2 and 5. In Eq. (2.15),q = q(T, p o , T o )
is a reference viscosity within the power law region, i.e., q at the reference
shear rate 7, at the reference pressure p o , and at the reference temperature
To.In the literature the power law model has been used very widely because
its form allows direct integration of the equation of motion for several flow
geometries to be carried out.
There are very few data on CI and p available; however, a0 and Po values
of the zero viscosity ro have been published for several molten polymers.
Thus, a relation will be derived in the following between CI and a, and
between p and Po. The viscosity curves measured at different levels of T
and p can be condensed to a single one, the so-called muster curue [31,32]

q(B3 7-9 p)/qo(T,P) = f (jJ * ?O(T P) 1. (2.16)


In the transformation process, the viscosity curves are moved in the log-log
plot in the direction of -45“; the shape of each curve remains the same.
Therefore, the shape of the master curve is identical with the shape of the
other viscosity curves. At larger values qoi) (> lo4 N m-’), ranges of the
master curve can be fitted again by the power law shown above

q ( k T, p)lqo(T, P) = w J q o ) ( l / m ) -l , (2.17)
and the viscosity becomes
~ ( 9 T,
, p ) = f+h/mj(l/m)-l. (2.18)
K is a “material constant” whose dimension depends on the value of the
power law exponent m ; K will be replaced by introducing the reference
viscosity Fj of Eq. (2.15). For certain ranges, the power law exponent m is
independent of temperature and pressure (since the master curve is indepen-
dent of temperature and pressure), but it slowly increases with higher values
of jq0 ranges. The pressure and temperature dependence of the viscosity q
is comprised by the pressure- and temperature-dependent zero viscosity qo
only. The pressure coefficient a,the temperature coefficient B, and the activa-
tion energy E of the viscosity in the power law region are then related to
21 8 HORSTH. WINTER

TABLE IV
ACTIVATION
ENERGYE , AND PRESSURE
COEFFICIENT
a,, OF
THE ZERO VISCOSITY
qoo

EO a0
Polymer [lo4 J/gm mole] [lo-’ m2 N-I] Literature

1.d.PE 5.44 -
1281
h.d.PE 7.08-8.99 3.25-3.98 C331
PS - 4.28-9.07 ~321
PMMA 14.23-19.1 3 2.45-4.08 1331

The temperature coefficient jo of the zero viscosity around


the temperature To is equal to E o / R T o 2 ,see Eq. (2.13). a and j
of the power law region can be determined from a. and Po by
dividing with m, see Eq. (2.19).

ao, Po, and Eo of the zero viscosity by

a = ao/m, P = Po/m, E = Eo/m. (2.19)


Table IV [28,32,33] lists E o and a. values of some polymers; the data can
be used to determine a and P of the power law which describes the viscosity
in the jqo range of the application in question. Semjonow [29] collected
Eo data from the literature which is quite extensive.
Due to the small values of a, the pressure dependence of the viscosity
usually can be neglected up to moderate pressure levels ( < 300 bar, de-
pending on the value of a) as they occur in extrusion. In injection molding
studies, however, the pressure may adopt values up to 1500 bar and the
pressure dependence should be included. In the following study, the pressure
dependence of viscosity is not taken into consideration, although the numerical
procedure would not have to be much different: it just would require one
or more iterations of the whole computation procedure, until the axial pres-
sure profile along the channel is known.
The concept of steady shear flow is a theoretical one and it can only be
approximated. However, the rheological properties of molten polymers do
not seem to be too sensitive to some deviation from steady shear flow; and
for a large number of applications, the results from steady shear flow cal-
culations agree with flow experiments reasonably well. The shear flow might
be unsteady (a/& # 0 ) ; during startups a constant stress is achieved only
after some time of development. But even if the flow is steady (a/& = 0), it
still might deviate from shear. Deviations occur as slowly relaxing entrance
effects; temperature changes along stream lines, which induce changes in
shear rate, are in contradiction to “steady shear flow,” which is defined to
be isothermal and at constant shear rate; in Poiseuille flow, pressure changes
VISCOUS DISSIPATION
IN FLOWING
MOLTENPOLYMERS 219

actually influence the density, while the fluid supposedly is incompressible.


Flows of this kind are called nearly steady shearjlows, where the adverb
nearly may refer to the word steady or the word shear. These limitations
of the applicability of the shear flow concept will be mentioned again, when
all the assumptions are listed together with the system of equations, see
Sections II.B.l and II.C.l.

Shear Flow Geometries with Open or with Closed Stream Lines


The most important shear flow geometries are shown in Fig. 3; Table V
[34-951 lists heat transfer studies on those flow geometries. The flow due
to the relative motion of one of the surfaces is called Couettejlow, while

Couette flow

Poiseuille flow

‘fl
Couette flow and Poiseuille flow superimposed

@ _..
v g
, \

0 0
- positive pressure gradient
____ negative pressure gradient

FIG.3. The main simple shear flow geometries [2]: (a) drag flow in the narrow slit between
two parallel plates (plane Couette flow), no pressure gradient; (b) axial drag flow between two
coaxial cylinders (annular Couette flow), no pressure gradient; (c) flow through a pipe with
constant circular cross section (Poiseuille flow); (d) flow through a narrow slit (Poiseuille flow);
(e) axial flow through an annulus (Poiseuille flow); (f) helical flow (flow through an annulus
with rotating inner cylinder); (9) axial drag flow in an annulus with nonzero axial pressure
gradient; (h) drag flow in the narrow slit between two parallel plates with nonzero pressure
gradient; (i) angular drag flow in the annulus between two coaxial cylinders (circular Couette
flow), no pressure gradient; (k) flow in a cone-and-plate or in a plate-and-plate viscometer.
Geometry a will be referred to as a, if the stream lines are open, or a2 if the stream lines are
closed (limiting case K + 1 of geometry i).
TABLE V : HEAT
TRANSFER SHEARFLOW
IN STEADY
STUDIES

Thermal Flow geometries with


Temperature boundary Flow geometries with open stream lines closed stream lines
field condition a1 b C d e f g h a2 i k

Fully developed constant see a, 34-41 15. 37. 42 43 31, 38,42, 38, 42, 45, 52, 53E,
wall 44-48 49-51E 54
temperature
constant
heat flux
at wall,
adiabatic
wall
mixed see a2 55E 42 38, 42. 45, 38,4549, 54
46, 55E. 51E,55E,
56,57E 58E, 59E
Developing constant 60.61 62-64.65E. 17,80-82 77, 83 82, 84, 86E, 87,88
wall 66,67E, 85
temperature 68E,
69-71,
72E, 73,
74,75E,
76-79,
133
constant 63,67E, 83 89
heat flux 68E, 70,
at wall, 71, 72E,
adiabatic 76, 78, 79

mixed 90E, 93E. 17,90E, 77,90E, 84.85, 89,92


94E, 95E 91E

Letters u, b, . . . ,k refer to Fig. 3, where flow geometry a occurs with open and with closed stream lines. The numbers refer to the list of literature.
Experimental studies or studies which contain experiments are marked with an E.
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 22 1

the flow due to a pressure gradient is called Poiseuilleflflow.In the literature,


most emphasis has been laid on the fully developed temperature field in
pipe flow and in Couette flow, and on the developing temperature field in
pipes with circular cross section. Experimental studies or studies that con-
tain an experimental part are marked by an E. Poiseuille flow in a pipe
with constant but irregular cross section or Poiseuille flow in curved chan-
nels with constant cross section induce some small secondary flow in the
cross section [96]; this will be mentioned in Section I11 on nonviscometric
flow.
A detailed description of the historical development of shear flow analysis
can be found in the introduction of original papers on the different prob-
lems (see for instance [55] for the fully developed temperature profiles and
[77, 781 for developing temperature profiles in pipe flow). This study tries
to describe the various aspects of shear flow analysis and give credit to the
different authors in connection with the arguments in the analysis.
Stream lines are lines whose tangents are everywhere parallel to the ve-
locity vectors. In steady flow, the stream lines describe the paths of fluid
elements. The heat transfer in the various shear flow geometries depends on
whether the stream lines are open (type a,-h in Fig. 3) or closed (type a,,
i, k in Fig. 3). In steady flows with open stream lines, the temperature is
locally constant with time (dT/dt = 0); for displacements along the flow
direction, however, it changes until a fully developed temperature field
(DT/Dr = 0 for T , = const) is reached, where conduction and viscous dis-
sipation balance. In processing equipment, the fully developed temperature
field is achieved rarely since the flow channels are not long enough and the
thermal boundary conditions usually change in the flow direction. Never-
theless, the calculated fully developed temperature field is very useful as a
reference state. The degree of development can be estimated from the value
of the Graetz number. The unsteady developing temperature ( d T / d r # 0,
aT/dz # 0) in flows with open stream lines has been studied very little,
possibly because the numerical or experimental techniques are very involved.
In shear flows with closed stream lines, the temperature is assumed to be
uniform along the stream lines, but locally changing with time (dT/a@= 0;
dT/dt # 0) during the starting phase. The stream lines are supposed to be
circles, and 0 is the coordinate in the flow direction. Convective heat
transfer has no influence on the temperature field. After some developing
time, a constant temperature field is reached where viscous dissipation and
conduction balance. The degree of development can be estimated from the
value of the Fourier number.
Drag flow in a narrow slit (plane Couette flow) is introduced twice (al
and a, in Table V). One might treat it as an entrance value problem (as in
a,) and study the axial development of the temperature field beginning from
222 HOWTH. WINTER

some inlet temperature distribution. On the other hand, one might treat
plane Couette flow as a limiting case of circular Couette flow (as in a2);
i.e., the stream lines are thought to be closed, and there are no changes in
flow direction (d/d@ = 0). The temperature develops with time, beginning
with some initial temperature distribution. The fully developed temperature
field is the same for both cases.

A. THERMAL
BOUNDARY
CONDITION
The specific heat flux 4 at the boundary is given by the thermal conduc-
tivity kRuidof the fluid together with the temperature gradient (aT/dr), in
the fluid layer next to the wall
4 = -ktluici(JT/ar)w* (2.20)
The problems of this section are described in cylindrical coordinates z, r, 0,
where I is the coordinate perpendicular to the wall.
If the thermal boundaries are not taken to be isothermal ( T , # const),
the thermal development in the fluid is connected with the thermal develop-
ment in the wall. The heat flux at the boundary is determined not only by
the conduction to the outside of the channel, but also by the thermal
capacity of the wall. Both effects will be analyzed separately in the following.

1. Biot Number and Conduction to Surroundings


If the effect of energy storage in the wall is of no influence, the heat flux
at the boundary generally depends on the difference of the temperature
level of the experiment to some temperature of the surroundings. In the
analysis, the temperature gradient in the fluid layer next to the wall is taken
to be proportional to the outer temperature difference (T, - T,); T, is the
temperature of the surroundings, and T , is the wall temperature, i.e., the
temperature at the boundary between melt and containing wall. The coeffi-
cient of proportionality is the Biot number [69,71] of equation
(dT/Jr), = Bi(T, - TJh. (2.21)
Bi is already well known for describing the thermal boundary condition
during the heating or cooling of solid bodies (see for instance [97,98]). h is
a characteristic length of the flow channel, i.e., the gap width of a slit or
the radius of a pipe.
Equation (2.21) describes just the radial heat flux in the wall; the axial
heat conduction in the wall is neglected in the Biot number. For shear flow
applications with closed stream lines, the validity of this assumption has to
be verified in each case. However, the assumption seems to be reasonable
VISCOUSDISSIPATION MOLTENPOLYMERS
IN FLOWING 223

for shear flow applications with open stream lines, where the axial tempera-
ture gradient is much smaller than the radial one; during thermal develop-
ment, the heat flux into the wall changes in the flow direction, but beyond
a certain distance from entrance into the channel, these changes are small
due to a small axial gradient.
The value of Bi for certain applications can be derived from a heat
balance for the wall. I n some cases Bi is a function of geometry and of the
thermal conductivities only. The heat flux through the wall of a pipe, for
instance, can be determined from the inner radius rp of the pipe wall, the
wall thickness s, the thermal conductivity kwall,and the inner and outer
temperatures T , and T , [98, p. 711.

(2.22)

On the other hand, the heat flux into the wall is determined by the thermal
boundary condition (Eqs. (2.20) and (2.21))

(2.23)

Thus, for steady pipe flow with controlled temperature at the outer wall
( T I = T w ;T , = Ts), the Biot number can be calculated by equating Eqs.
(2.22) and (2.23):
(2.24)

Applying this formula to capillary viscometry (rp= 0.5 cm, s = 4.5 cm), one
finds values of Bi w 20. Examples for pipe flow with 1 I Bi I 100 are given
in Fig. 8 of Section II.B.4. Similarly, the outer Biot number Bi, for annular
flow (with ri and r, as the inner and outer radius) would be

(2.25)

and some examples for pipe extrusion give 1 < Bi, < 10.
Figure 4 illustrates the geometrical meaning of Bi. The tangent to the
temperature curve T(r/h)at the wall passes through a guide point outside
the flow channel; the distance between the guide point and the wall is Bi-',
and the ordinate is the surrounding temperature T,. For Bi = 10, for in-
stance, the distance of the guide point from the boundary is 1/10 of the gap
width h for annular flow or 1/10 of the radius r, for pipe flow. When the
Biot number changes in flow direction, one can visualize this by the appro-
priate displacement of the guide point.
224 HORSTH.WINTER

FIG.4. Thermal boundary condition for channel flow described by the Biot number Bi and
the surrounding temperature T,, i.e., by a guide point outside the channel.

The boundary condition for the temperature field is not known in general.
If there are temperature data T,(z) available, they can be used in a numerical
program. But often one has to guess these conditions to make an estimate
of the temperature profiles possible. Most of the studies shown in Table V
prescribe idealized conditions such as:
constant wall temperature
T, = const or Bi, -, - co, Bi, + 00; (guide point at the wall),
constant heat flux at wall
(dT/dr), = const or Bi(T, - T,) = const,
adiabatic wall
(aT/dr), = 0 or Bi = 0; (guide point at infinity).
The use of the Biot number allows one to adopt more realistic thermal
boundary conditions, and one goal of further experimental heat transfer
studies should be the measurement of Bi in various engineering applications.
For the examples shown throughout this analysis, the boundary condition
at the wall will be described b y Bi and T, independent of z (for steady flow
with open stream lines) or independent oft (for flow with closed stream lines),
respectively. If the value of Bi is finite, the wall temperature T,(z) or T,(t)
changes according to the development of the temperature field, and it reaches
a constant value Tw,m in the fully developed temperature field.
DISSIPATION
VISCOUS IN FLOWING
MOLTENPOLYMERS 225

2. Thermal Capacity of the Wall


The Biot number is appropriate for describing heat conduction to the
surroundings. However, if the wall stores some energy during thermal de-
velopment, a different boundary condition is needed to describe this effect.
The thermal development should be calculated for the fluid and the wall
together. This has been done by Powell and Middleman [92] for plane
Couette flow with one wall having a finite mass, which absorbs part of the
heat generated by viscous dissipation. The thermal development was found
to be significantly retarded by the response of the boundary; the parameter
characterizing the retardation is the ratio of mass times heat capacity of the
solid wall and that of the fluid: (mcp)wa,,/(mcp)R,,d.
In this study, however, detailed calculation of the temperature field in the
wall will be-avoided by introducing a capacitance parameter C . For flow
with closed stream lines, the wall temperature changes with time during the
thermal development. The rate of thermal energy stored in the wall is
assumed to be proportional to the time change DT,/Dt of the temperature
at the boundary. The temperature gradient in the fluid layer near the wall
becomes

(2.26)

The capacitance parameter C is dimensionless, and the ratio hla,,, is used


in Eq. (2.26) because below the whole boundary condition will be made
dimensionless. C is determined by the geometry and by the capacitance of
both the fluid and the wall. The heat flux to the surroundings is kept propor-
tional to the outer temperature difference T, - T,.
Assuming constant thermal properties of the wall material, the rate of
energy storage in the wall is proportional to the time change of the average
temperature of the whole wall. The time change of the average temperature
of the wall might not be proportional to the time change DT,/Dt at the
boundary. Thus, the capacitance parameter describes the effect of energy
storage in the wall only approximately. In many polymer engineering appli-
cations, however, the thermal development in the wall is much faster than
the thermal development in the fluid, and the temperature at the boundary
is representative for the whole wall. An example where uniform temperature
in the wall is assumed will be given in the following.
Example for C: The temperature of the inner cylinder of a Couette sys-
tem (geometry i in Fig. 3, ri = inner radius, r, = outer radius, h = r, - r i )
changes during the thermal development of a shear experiment. The tem-
perature of the inner cylinder is assumed to be uniform and equal to the
226 HORSTH. WINTER

temperature at the boundary; this is justified, if the ratio of the Fourier


numbers (see Section II.C.2) is small:

(2.27)

Axial heat conduction is neglected and, of course, the Biot number is equal
to zero. The heat flux into the inner cylinder is balanced by the temperature
raise of the inner cylinder:
2nri(aT/ar)w kRuid = nri2(PC)cylindcr aTw/at. (2.28)
By comparing Eq. (2.28) with Eq.(2.26)one finds the capacitance parameter
for the Couette system:

(2.29)

The influence of C on the thermaI development will be shown in Figs. 23


and 24 of Section II.C.3.
For most of the steady heat transfer problems with open streamlines, the
walls are stationary or just rotating about the z axis. The capacitance of
the wall has no influence on the temperature (DT,,,/Dt = 0). An exception
would be the inner boundary of axial Couette flow (geometry b or g of

);r
Fig. 3) as occurs in the wire coating die. The corresponding thermal bound-
ary condition is
Ts - Tw + ci--. vwh aTw
= Bi, (2.30)
W h athid aZ

u, is the axial velocity of the wall (inner cylinder).


For the example of a wire coating process, a heat balance for the inner
cylinder (wire) leads to the same formula for the capacitance parameter as
for the Couette system above, Eq. (2.29). The assumptions made were uni-
form temperature in cross section of wire and no axial conduction in the
wire. The Biot number for the wire is zero.
The geometrical meaning of the boundary condition with capacity and
conduction to the surroundings is shown in Fig. 5. The guide point is not
at a constant position (as for flow with DT,/Dt = 0; see Fig. 4),but moves
during the thermal development along T, = const. For the fully developed
temperature field, the wall temperature does not change any more; the
capacitance of the wall is of no influence, and the guide point is, as in
Fig. 4, at a distance Bi- from the boundary.
An adiabatic or perfectly insulating wall would be a wall without thermal
capacitance (or with an appropriate heat source of its own); the corre-
sponding Biot number and the capacitance parameter are both equal to zero.
VISCOUSDISSIPATION MOLTENPOLYMERS
IN FLOWING 227

displacement of guide point


during thermal development

FIG.5. Thermal boundary condition for a wall with thermal capacitance.During the thermal
development,the guide point of the tangent on the temperature field moves toward the position
’,
(Bi- T J .

B. STEADYSHEARFLOWWITH OPENSTREAMLINES
Steady shear flow with open stream lines could be analyzed now by
going into each of the shear flow geometries al-h in Fig. 3. Instead, it will
be demonstrated here that the helical flow geometry is representative since
all the other geometries are limiting cases of helical flow.
In helical flow, the fluid flows through an annulus between two con-
centric cylinders (Fig. 6). Axially, the fluid flows due to a pressure gradient
and/or due to the axial movement of the inner (or outer) cylinder. In the
circumferential direction, the fluid flows due to the rotation of the inner
cylinder. Fluid elements move on helical paths; the angle of the helices

flow due to pressure gradient flow due to rotation of


/and due to axial movement inner cylinder,

path of
!bid particle
/
/
/
1

//// / / /// / / / /’/ ’,’,.’//,’/,’//// / / / / ’ i//

FIG.6. Helical flow geometry.


228 HORSTH.WINTER

depends on the ratio of the axial to the circumferential velocity component,


which both depend on the radial position of the fluid element.
The annular geometry is characterized by the ratio of the radii
IC = ri/ra, (2.31)
The limiting cases are the pipe ( I C = 0) and the plane slit ( K -, 1). The posi-
tion in the annulus is given by dimensionless coordinates :
r - ri
radially: Y = ___ , O I Y I 1 ; (2.32)
ra - ri
Z
axially: 2 = 0 I Z I Gz-’. (2.33)
1 Gz’
~

The value of Z indicates, as will be shown in the following, to what degree


the temperature field is developed along the channel. The Graetz number
Gz will be defined in Eq. (2.56).
1. Assumptions and System of Equations
The equations of change, Eqs. (1.1)-(1.3), have to be simplified before they
can be solved. First the assumptions will be listed, then they will be com-
mented upon:
incompressible fluid with constant thermal conductivity and diffusivity ;
steady laminar flow (a/& = 0);
rotational symmetry (a/a@ = 0 ) ;
velocity gradients

no slip at walls;
inertia negligible ; kinematically developed velocity at z = 0;
gravity negligible;
viscosity measured at constant temperatures and constant shear rates
gives applicable local values of the viscosity during temperature changes
and during small changes in shear rate; rheologically developed stress at
z = 0;
convective heat transfer much larger than conduction in flow direction;
heat transport toward the walls by conduction only.
Throughout this section the molten polymer is taken to have constant den-
sity ( p / p = 1; E = 0), constant thermal conductivity (k/E = l), and constant
thermal diffusivity (a/a = 1). In the general system of equations for helical
flow, however, these properties have been kept as variables, and one might
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 229

evaluate them in the numerical program using p ( p , T ) , Up, T ) ,and u(p, T )


data from measurements on the fluid at rest.
The density is assumed to be constant since in actual experiments (with
T # const, p # const) density changes are delayed, i.e., the changes are
overestimated if one applies p data of equilibrium thermodynamics. For
this reason, the effect of expansion cooling is not considered in all but one
example. For the one exception, the supposed expansion cooling term
(containing E T )of Eq. (1.8) is kept in brackets in the energy equation; the
effect of expansion cooling is estimated in an example of pipe flow; see
Fig. 13 with E # 0.
During the axial development of the temperature field, the temperature-
dependent viscosity is changing and causes the shape of the velocity profiles
to change accordingly. For continuity reasons, the changes of the axial
velocity require some radial flow. Using the equation of continuity, Eq. (l.l),
the radial velocity components have been estimated from the change of the
axial velocity component and have been found to be small (lo,( -=lo-’ ij).
Throughout Section ILB, the influence of the radial velocity components on
the radial heat transfer, on dissipation, and on the viscosity will therefore
be neglected.
Due to rotational symmetry and due to du,/dr and r a(u,/r)/dr being the
largest gradients, the isotropic pressure is taken to be a function of z only:
P = P(4.
The shear rate becomes

(2.34)

which is the root of the second invariant of the rate of strain tensor.
In most applications, molten polymers do not slip at the wall, and in all
the published heat transfer studies this assumption has been made. Polymeric
materials such as high density polyethylene, polyvinylchloride, or poly-
butadiene, however, seem to slip in certain ranges of the normal stress and
shear stress at the wall [99,100]; the velocity field is then drastically changed
and additional frictional heating occurs on the sliding surfaces.
The velocity at the entrance ( z = 0) is assumed to be fully developed;
i.e., inertial effects are neglected, and the stress is assumed to be governed
by the three viscometric functions (of steady unidirectional shear flow) at
the local shear rate and the local temperature. For low Reynolds number
pipe flow of inelastic liquids, the kinematic development is practically com-
pleted after a length of 1 = O.lra Re [loll. Neglecting inertia might there-
fore be justified for entrance flow of molten polymers, which is low Reynolds
number flow.
230 HORSTH. WINTER

The rheological properties of the polymer entering the annulus are deter-
mined by a flow and temperature history, which obviously is different from
that for steady shear flow. Judging from the measured pressure profiles
along a slit die, steady shear flow might be reached practically at l/h = 20-30
(depending on geometry, flow rate, and polymer melt). Therefore, the heat
transfer study of this section may give unrealistic results for flow in short
annuli and short circular holes, which rheologically should be treated as an
entrance flow problem.
The equations for conservation of mass, momentum, and energy in helical
flow are

(2.35)

The average axial velocity is


- 2
u, = u,r dr. (2.36)
ra2 - riz
The initial and the boundary conditions are

T(r,0) = T,(r)

(2.37)
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 23 1

The meanings of Bi and T , have been described already in Section 1I.A on


the thermal boundary condition (Bi, <c 0; Bi, > 0). The capacitance param-
eter C of Eq. (2.30)has been omitted here since the temperatures are assumed
to be steady and since the walls are stationary for most shear flow applications
with open stream lines.
For the dimensionless presentation of the equations, a reference velocity
B can be defined by vector addition of the mean axial velocity b, and the
mean circumferential velocity ~ , , ~ / 2 :
-
v = [vz2 + (VeJ2) 2 ] 112 . (2.38)
The reference length is taken to be the gap width
h = r, - r i . (2.39)
For pipe flow the reference length becomes equal to the pipe radius (h = ra).
Using the reference velocity 3 and the reference length h, one can define a
reference shear rate -
p = b/h (2.40)
and a reference viscosity
-
rl = v(74 To). (2.41)
T o is a characteristic temperature level of the experiment, for instance the
average melt temperature at the inlet ( T o= Te).Flow problems with viscous
dissipation do not have a characteristic temperature difference to
which temperature changes can be related. Some authors relate the tempera-
ture to the temperature level T o ;this, however, seems to be rather arbitrary
since T / T o might assume different values in otherwise similar processes
(e.g., at different temperature levels T oand To').The value of T / T oaddition-
ally depends on the choice of temperature scale. Therefore, the temperature
coefficient of the most temperature-sensitive property, the viscosity, has
been used to define the dimensionless temperature: (AT)ref= 0-l.
The dimensionless variables are:
velocity (vR> v@, = ( v ~ / u@/B,
~ , uz/D)? (2.42)

pressure gradient (2.43)

shear stress (2.44)

(2.45)

radial position
(2.46)
232 HORSTH . WINTER

viscosity

temperature

The dimensionless form of the equations becomes

(2.49)

(2.50)

(2.51 )

The dimensionless average axial velocity is

VZ =
2
1 VzR dR, (2.53)

(2.54)
VISCOUS DISSIPATION MOLTENPOLYMERS
IN FLOWING 233

In the expansion cooling term of the energy equation, the absolute tempera-
ture T is not replaced by the dimensionless temperature 9 since the dimen-
sionless product ET can be considered as constant within the accuracy of
the calculations.
The system of equations is formulated in cylindrical coordinates, and
R = r/ra is kept as a dimensionless coordinate, even if h = r, - ri is the
reference length and not rs. The results are presented using the dimension-
less coordinate Y . If one wants to avoid this inconsistency, the substitution
R = Y(l - K) + K, dR = dY(1 - K)

eliminates R from the equations; with this substitution the equations look
unnecessarily complicated and therefore both coordinates are kept : R in the
equations and Y in the graphical presentation of the results. For pipe flow,
R and Y are identical.

2. Dimensionless Parameters
The problem as stated in Eqs. (2.49)-(2.54) is completely determined by
six dimensionlessparameters (Na, Gz, K , m, Vz,L ) together with the boundary
conditions. A general description of the dimensionless parameters has been
given by Pearson [3]. If the pressure dependence of the viscosity or non-
constant thermal properties would be included, the number of parameters
would increase accordingly.
The equation of motion and the equation of energy, Eqs. (2.35),are coupled
by the temperature-dependent viscosity. The extent of the coupling increases
with the value of the Nahme number [44] :
Na = j02q/X (2.55)
which compares the dissipation term with the conduction term in the
equation of energy. For values of Na greater than 0.1-0.5 (depending on
geometry and thermal boundary conditions), the viscous dissipation leads
to significant viscosity changes, i.e., changes reflected in the T and u fields.
For smaller values of Na, isothermal conditions can be achieved practically;
in this case, the equation of motion can be integrated independently of the
energy equation.
In some studies the Brinkman number [62] Br = B2ij/kTo has been used
instead of the Nahme number. However, Br contains the arbitrary tempera-
ture level To(since no characteristic temperature difference is available) and
may, therefore, have very different values for similar processes. The value of
Br does not give any information on the extent of the coupling between the
equation of motion and the energy equation. (Note that the Nahme number
sometimes is called the GrBith number after Griffith [102], who used the
same dimensionless group in one of the later applications.)
234 HORSTH. WINTER

The energy equation contains a convection, a conduction, and a dissipa-


tion term. By comparing the convection and the conduction terms one
arrives at the Graetz number [1031
GZ = Vzh2Jzil (2.56)

which has been included in the dimensionless form of the z coordinate. The
Graetz number can be understood to be the ratio of the time required for
heat conduction from the center of the channel to the wall and the average
residence time in the channel [75]. A large value of Gz means that heat
convection in flow direction is more important than conduction toward the
walls, Gz = 100, for instance, is a common value for extruder dies. (Note
that some authors define the Graetz number Gz n.)
The Gz number has been defined with the average axial velocity and the
length of the annulus, instead of the reference velocity ij and some mean
path length Tfor the fluid elements in the annular section. One might, how-
ever, define Tto be T = 1ijpzwhich would result in a Graetz number Gz =
ijh2/uTequal to the one defined in Eq. (2.56).
The value of the dimensionless average axial velocity

Y.=B=[I+(gJ]
V
2 -1/2
, O<Vz<1, (2.57)

describes whether the flow tends to be closer to axial flow in an annulus


(Vz= 1) or closer to circular Couette flow (Vz= 0).
The dimensionless length of the annulus is
L = l/h = l/(ra - Ti). (2.58)
Another dimensionless parameter originates from the shear dependence of
the viscosity: the power law exponent m in Eq. (2.47).

3. Universal Numerical Shear Flow Program


The system of equations is solved by an iterative implicit method (aJaZ
described by a backward difference; alaR and a2faR2 described by center
differences;gradients at a boundary are calculated from a parabola through
three points), similar to the one used in an earlier study on helical flow [85].
A network is superimposed on the annulus. Difference equations are then
derived for each node point, which fulfill the condition of conservations of
mass, momentum, and energy. The method described in the following was
found to converge rapidly; for example, a run with 60 radial and 250 axial
steps requires a computation time of about 30 s.
The solution procedure is an iterative one, in which the coupled equations
are linearized and solved separately. The nonlinear terms and the coupling
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 235

conditions have to be satisfied by alternating improvements on the velocity


and on the temperature field. The iteration is terminated when the rela-
tive change in successive steps becomes smaller than one thousandth
( A ~<~ ,1 0 - 3 ) .
The flow chart in Fig. 7 describes the structure of the program. The velocity
field at the enrrunce, which is assumed to be fully developed kinematically,
is calculated by taking, as a first guess, a Newtonian viscosity according to
the entrance temperature field. The shear dependence of viscosity is included
then by iteration, using the improved values of the velocity field. After about
6-20 iterations, the velocities reach values that are practically constant.

geometry. material and process data, boundary conditions


.
Bi, , Bi., G.,, c,o ve,, , v,,, ; entrance conditions T (r);
z=o; t,=l, v e z v,= 0 .

dimensionless parameters No , Gz,Vz ,x , dR = const.


logarithmic steps sizes AZ, , number of steps no , n -- 0
I

conditions ; 4,. 0

end

FIG.7. Flow chart of universal shear flow program.


236 HORSTH. WINTER

The axial velocity V, and the pressure gradient P' are calculated from the
Z component of the equation of motion together with the integral of the
flow rate (Eqs. (2.51) and (2.53)). The circular flow V, is evaluated from the
0 component of the equation of motion (Eq. (2.50)).The R component V,
of the velocity supposedly does not influence the viscosity and the convec-
tion; thus it is calculated separately at the end of each step using the equation
of continuity (a numerical method that allows for positive or negative radial
flow contributions has been suggested by Gosman et al. [104]).
The entrance conditions are then stored, and the fully developed tempera-
ture jield is calculated so as to be available as a reference state for the
developing temperature field. In the fully developed temperature field, the
convective term of the energy equation is zero. The iteration starts out with
the viscosities and velocities at the entrance. They give a first approximation
of the dissipation term and of the fully developed temperature field. Using
this solution, one gets improved values of the viscosities and the velocities
by iteration. These values of the viscosities and velocities lead to the second
approximation of the fully developed temperature field, and so on. After
satisfying the condition of (A,,J < the values of (9, - ~ ( I c00))
, and
of (9, - 9(1, 00)) are stored as reference values for the developing tempera-
ture field. Then the program goes back to the entrance temperature field
and starts calculating the developing temperature, velocity, shear stress, and
pressure. If both walls are adiabatic (Bi, = Bi, = 0), there does not exist a
fully developed temperature field; the program then starts calculating the
developing temperatures immediately.
For flow in a capillary (IC= 0), the velocity and the temperature have a
zero gradient at R = 0.
The power law model fails in describing the viscosity at low shear rates,
and for computational purposes at least one has to set an upper limiting
value of the viscosity (this has been done in the numerical program of this
study). For more accurate calculations, one has to approximate ranges of
the viscosity curve by several power law and temperature coefficients. Also
a viscosity table could be used instead.
The numerical program has been checked with analytical solutions of the
fully developed temperature and velocity field in plane Couette flow and
with isothermal flow in a pipe and in an annulus [lOS].
Besides helical flow with its steady, but developing temperature field, the
system of equations, Eqs. (2.35)-(2.37), describes the flow geometries of all
other steady shear flows with open stream lines (type a,-h in Table V).
Therefore, it actually is possible to use one numerical program for all these
flow cases. The appropriate values of the ratio of the radii IC,the axial velocity
of the inner cylinder V,(IC,Z ) , and the average axial velocity Y' are listed in
Table VI.
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 237
TABLE VI
SHEAR AS LIMITING
FLOW GEOMETRIES CASES
OF HELICAL
FLOW"

Flow geometry as
described in Fig. 3 u V.(JG z) VZ

a1 0.999 2 1
b o<u<1 determined by 1
iteration for P = 0
0 - 1
0.999 0 1
O < X < l 0 1
O<S<l 0 O<Vztl
O<S<l finite 1
0.999 finite 1

a2 0.999 0 0
1 O<W<l 0 0

Listing of the corresponding geometry (described with K ) and kinematics


(described with the velocity of the inner cylinder I/Z(K.Z)and the average
axial velocity Vz).Geometries a,-h are with open, and geometries a2 and i
with closed stream lines.

Due to the parabolical character of the solution procedure for the equation
of energy, the helical flow program can be applied only to flows with non-
negative velocity components. Thus, axial drag flow in an annulus with
nonzero axial pressure gradient (type g) and drag flow in a narrow slit
between two parallel plates with nonzero pressure gradient (type h) can be
analyzed only up to moderate positive pressure gradients. A solution proce-
dure that allows for back flow is described in the literature [104], but it
does not seem to have been applied to these types of flow.

4. Calculated Results
There is a large variety ofheat transfer problems solvable with the universal
shear flow program. Some examples follow, mainly concerning the thermal
boundary conditions (Biot number) and the kinematics for various shear
flow geometries. Similar examples of helical flow or annular flow calculations
have already been published, however, with idealized thermal boundary
conditions [85]. In all the examples of this section, the entrance temperature
(at 2 = 0) is taken to be 9,(R) = 0.
The thermal boundary conditions influence the developing temperatures
and velocities to a large extent. In analytical studies generally, idealized
conditions are assumed, i.e., isothermal or adiabatical wall; in real flow
238 HORSTH. WINTER
I 2
No.5
Bi =lo0
=m rn =2.5 rn = 2 . 5
31R.Z

A
1

.?=lo-’
z 40”

C
0.5 R a5 R 1

0,5- R 1

Ro.8. Influence of the thermal boundary condition, described by a guide point outside the
channel given by Bi and 9, = 0, on the developing temperature field in pipe flow. Bi = 100 is
close to the isothermal wall condition, while Bi = 1causeslarge changes of the wall temperature.
VISCOUS bISSIPATION IN FLOWING
MOLTENPOLYMERS 239

situations the thermal boundary condition is somewhere between the two.


The strength of the Riot number in describing a more realistic kind of
boundary condition will be demonstrated on pipe flow: the fluid supposedly
enters the pipe at a constant temperature equal to the temperature of the
surroundings (9, = 9, = 0).
Due to viscous Pissipation, the temperatures increase in flow direction
(Fig. 8). For Bi = 100, the wall temperature stays nearly constant, while for
Bi = 10 and Bi = 1, the wall temperatures already increase in early develop-
ment. The fully developed temperature at Bi = 10 has a value between that
at Bi = 100 and Bi = 1. At large Bi the temperature in the layer near the
wall is kept low; the viscosity and hence the viscous dissipation is large, and
a large temperature gradient is needed to conduct away all newly dissipated
energy. At small Bi the wall temperatures have to rise signifieantly before the
heat flux at the wall can balance dissipation: the temperature gradient can
still be relatively small since the viscosity and hence the viscous dissipation
become small at high temperatures. At large Bi the temperathreh are high
because viscous dissipation is most pronounced. At small Bi the temperatures
are high because the conduction toward the surroundings requires large wall
temperatures. For intermediate Bi, the fully developed temperature has a
minimum.
The corresponding pressure gradient P ( Z ) decreases due to the decrease
of the temperature dependent viscosity (Fig. 9). The decrease of P'(Z1 below
its value at the entrance P(0) is most pronounced at small values of the
Biot number, where the wall temperatures increase the most.
The temperature gradient at the wall is defined with the Biot number and
an outer temperature difference 9, - 9, (see Eq. (2.54)),where $,(Z) itself
depends, besides the other parameters, on Bi. The dimensionlesstemperature

FIG.9. Pressure gradient in pipe flow at different thermal boundary conditions; decrease due
to the thermal development described in the previous figure. The pressure gradient for isothermal
pipe flow can be calculated analytically: P(0) = - 2 ( m + 3)"".
240 HORSTH. WINTER

gradient [as(&2)/dRlw increases with Z till it reaches its fully developed


value at about Z = 1 (see Fig. 8).
a. Nusselt Number. For engineering calculations, the specific heat flux q
often is described by means of the Nusselt number [lo61

(2.59)

and a characteristic temperature difference (AT)ref.The product Nu k/h


sometimes is called the heat transfer coefficient. The value of the Nusselt
number depends much on the choice of the temperature difference ( A T ) r e f .
For flow without signijcant viscous dissipation, one takes it to be the local
average temperature difference ( T , - T ( Z ) ) or the average temperature
change ( T ( Z ) - T(0))in the flow direction. The average temperature of the
melt is chosen to be the “cup mixing temperature”

T ( Z ) = ___ T ( R , Z)Vz(R, Z ) R dR (2.60)


1 - uz
which would be the temperature of the homogeneous fluid after mixing
(constant specific heat per volume assumed).
The Nusselt number in its usual definition is not adequute for describing the
wall heat flux in flows with signijicant viscous dissipation. One disadvantage
of the use of Nu is the fact that both Nu@) and T ( Z )have to be known to
calculate the wall heat flux q ( Z ) ;the main disadvantage, however, is that in
Nu an attempt is made to describe two fairly unrelated quantities as a
function of each other, the temperature gradient at the wall and the average
temperature difference (Tw(Z)- T ( Z ) ) .This will be explained in the fol-
lowing, using pipe flow as an example.
Figure 10 shows developing temperature profiles in pipe flow, on the
left-hand side with negligible viscous dissipation (Na = 0.001) and on the
right-hand side with significant viscous dissipation (Na = 1). The wall
temperature is above the entrance temperature (9, = 0; 9, = 0.1); developing
temperatures at constant wall temperature (Bi = 00) are drawn as solid lines,
while for a thermal boundary condition with Bi = 10 dashed lines are.used.
For flow without signijcant dissipation (Na = 0.001), the temperature of the
fluid gradually approaches the wall temperature; the temperature gradient
at the wall (shown in Fig. 11) decreases monotonically till it becomes zero.
However, if there is significant viscous dissipation (Na = l),the temperature
in the layer next to the wall increases drastically even before the average
temperature is changed much. The temperature gradient changes its sign;
the fluid heats the wall, even if the average fluid temperature is below the
VISCOUS DISSIPATION
IN FLOWING
MOLTENPOLYMERS 24 1

9IR.ZI

0.: ~
0.2
----- Bi =10
-Bi =w

S(R,
3, = 0.1

O! 01

z=10-’

C
0.5 R 1

FIG.10. Calculated temperatures in pipe flow with wall temperatures above the entrance
temperature. Comparison of the thermal development without (Na = 0.001) and with (Na = 1)
significant viscous dissipation. The wall temperature is taken to be constant (Bi = a)or
described by a guide point with Bi = 10.

wall temperature. This corresponds to a negative Nusselt number or a negative


heat transfer coefficient, which is an unrealistic result.
The dimensionless average temperature difference (Fig. 12)

S ( Z ) - 3,(2)= p ( T ( Z ) - T , ( Z ) ) = B A T (2.61)
is negative at the entrance of the pipe (prescribed initial condition) and at
least for Bi = co, it increases monotonically with Z . For Na = 0.001, the
fluid approaches the wall temperature (9, = LJ,+,); for Na = 1, the average
temperature difference goes through zero and approaches a constant value
greater than zero. The Nusselt number Nu(Z), which conventionally is
defined with this temperature difference, has a singularity when the average
242 HORSTH.WINTER

FIG.11. Wall temperature gradients for pipe flow with viscous dissipation (Na = 1) and
without (Na = 0.001).The wall temperature is above the entrance temperature.

FIG.12. Development of the difference between the average temperature and the wall
temperature in pipe flow with viscous dissipation (Na = 1) and without (Na = 0.001); the
wall temperature is above the entrance temperature.
VISCOUSDISSIPATION
IN FLOWINGMOLTEN
POLYMERS 243

temperature difference AT is equal to zero [69,76,82]. The specific heat flux


at the wall obviously is finite (see the wall temperature gradient in Fig. 11).
This seeming singularity (at Z of AT = 0) suggests that the usual choice of
the reference temperature difference A T cannot be applied meaningfully to
flow problems with viscous dissipation [69]. The same argument is valid if,
instead of a constant wall temperature, a guide point is chosen outside the
channel, see Figs. 10-12 with Bi = 10.
In the definition of the Nahme number (Eq. (2.55)), p-’ is taken to be a
characteristic temperature difference, and for defining a Nusselt number for
flow with viscous dissipation one similarly may take

(ATLeF = B-’. (2.62)


Introducing this reference temperature difference into Eq. (2.59),the Nusselt
number becomes
Nu = hg(dT/dr), = (ds/aY),, (2.63)
and it then is identical with the temperature gradient at the wall. Equation (2.63)
seems to be an adequate definition of Nu for flow with viscous dissipation.
The relation between Nu and Bi is given together with Eq. (2.54).For fluids
with practically temperature-independent properties, an adequate definition
of the Nusselt number for dissipative flow can be made by means of the
“recovery temperature” [98, p. 4171.
b. Expansion Cooling. In steady flow of compressible fluids with nonzero
pressure gradient, the density will change in the flow direction; the equation
of energy, Eq. (1.7),contains a term that describes the cooling or heating due
to those density changes
cT DplDt,
where r T can be determined from Eq. (1.10). In the example of Fig. 13 the
influence of expansion cooling is shown on developing temperature profiles
in pipe flow (with 9, = 9, = 0). Values of ET = 0, 0.1, 0.2, 0.3 have been
used since values of this magnitude can be evaluated from equilibrium
thermodynamic data of molten polymers at rest. The applicability of equilib-
rium data to regions of rapid pressure changes is still an open question.
c. Thermal Development. For constant inlet temperature equal to the
temperature of the surroundings, the average temperature increases during
the development of the temperature field with increasing Z. The temperature
is fully developed at 2 = 0.5-2. Although the absolute value of the average
temperature depends on the power law exponent m, or Na, and on Bi, the
relative development is nearly the same for the different examples of pipe
244 HORSTH. WINTER

i
0 0.5 R

FIG.13. Calculated temperature profiles in a pipe with constant wall temperature equal to
the entrance temperature. The magnitude of ET determines the amount of cooling due to
expansion with decreasing pressure p; T is the absolute temperature.

flow (Fig. 14).In annular flow, however, the relative thermal development
depends on the thermal boundary conditions (Fig. 15); if the dissipated heat
can be conducted to both walls (Bi, = -10'; Bi, = lo5), the developing
length is much smaller than if one wall is nearly adiabatical (Bi, = -1;
Bi, = lo5).For Newtonian fluids (m = l), the channel length required for
thermal development is shorter than for fluids with shear dependent viscosity
(m = 3; rn = 5, for instance).
d . Zero Pressure Gradient or Zero Wall Shear Stress. The versatility of the
program will be demonstrated on some velocity profiles of isothermal flow
in an annulus, including the limiting cases of a plane slit (K:+ 1) and pipe
(K = 0). Annular flow with zero velocity gradient at the inner wall (which
also means zero shear stress at the inner wall) can be achieved with the
appropriate pressure gradient and the appropriate velocity V'(K) at the inner
wall (Fig. 16). An application of this type of velocity field could be die flow
in the wire coating process: operating at low shear stress at the surface of
the wire prevents ruptures of the wire. If the pressure gradient in the annulus
DISSIPATION
VISCOUS IN FLOWING
MOLTEN
POLYMERS 245

I
10-2 10-1 1 10
axial position 7

1o-~ 1o-z 10” 1 10


axial position 7
FIG.14. Development of the average temperature in pipe flow at different Bi and m ;(a) at
Na = 1 and (b) at Na = 5.

axial position Z

FIG.15. Development of the average temperature in annular Poiseuille flow at different Bi,,
Na, and m. K = 0.5; Q = 0.
246 HOWTH. WINTER

9
I1

1.0 -
x
z
->aB
0

-._
0

2 0.5 -

0 1
0 as
radial position Y

FIG.16. Calculated velocities and pressure gradients for isothermal axial flow in an annulus
with zero shear stress at inner wall. Parameter is the ratio of radii K.

is prescribed to be zero (P' = 0), the axial velocity of the inner cylinder V'(K)
has to be the larger, the smaller IC is (Fig. 17); for a plane slit (K + l), the
velocity gradient is constant and the velocity of the moving wall is twice the
average velocity, obviously. Taking different values V,(K) in an annulus of
K = 0.4 (Fig. 18), the pressure gradient P' adopts positive or negative values.
A zero shear stress at the inner wall or a zero pressure gradient at isothermal
flow does not mean that this condition applies to the whole flow channel:
due to the thermal development the velocity changes, and accordingly a
nonzero shear stress at the inner wall or a nonzero pressure gradient arises.

5 . Experimental Studies
The main motivations for undertaking experimental studies on heat
transfer in steady shear flow with open stream lines seem to be:
investigatingthe validity of the assumptions made in the analytical studies;
information on the thermal boundary conditions, i.e., values of Bi in
different applications.
The flow geometries chosen for experiments were pipe flow and helical flow
(see Table V). The measurable quantities were the flow rate, the pressure
VISCOUSDISSIPATION MOLTEN
IN FLOWING POLYMERS 247

radial position 'k


FIG.17. Calculated isothermal velocity profiles in an annulus at zero pressure gradient.
Parameter is the ratio of radii K.

"0 0.5 1
radial position Y
FIG.18. Calculated isothermal velocity profiles in an annulus of K = 0.4. Depending on the
values Vz,i,the pressure gradient P' adopts positive or negative values.
248 HORSTH. WINTER

profile, the radial temperature distribution at the inlet and at the exit, the
thermal boundary conditions ; additionally, for helical flow one could mea-
sure the torque and the angular velocity of the cylinders. As input data for
the numerical program one needs the flow rate (or a pressure gradient), the
properties of the polymer (viscosity q(P, T ) ,thermal diffusivity a(T),thermal
conductivity k(T),density p(p, T ) ) ,the melt temperature at the inlet Te(r),
and two boundary conditions each for the temperature and the velocity
fields. The other data can be used for a check on the validity of the numerical
solution. Several of the published experimental studies do not specify the
data needed for comparing with analytical solutions.
While wall temperatures can be measured quite accurately, the temperature
measurements in the flowing molten polymer always contain some systematic
errors. A thermocouple mounted on the tip of a probe is placed into the melt
stream. The probe is supposed to adopt the melt temperature as closely as
possible (zero temperature gradient in the polymer layer next to the probe).
Apart from distorting the velocity profile by introducing the probe into the
flow, two effects are influencing the temperature measurement: heat conduc-
tion along the probe, which requires a heat flux and a temperature gradient
in the polymer layer next to the wall of the probe, and viscous dissipation in
the polymer around the probe.
The error due to conduction along the probe can be excluded by setting
the base temperature, where the probe is mounted to the wall of the channel,
equal to the temperature at the tip of the probe [107,108]. The error due to
dissipation cannot be avoided, but it can be kept small by measuring the
melt temperature at positions of very low velocities, i.e., after slowing down
the flow in a wide channel and then calculating back to the corresponding
temperatures at the exit of the narrow channel by means of the stream func-
tion [91,109]. Van Leeuwen [110] studied the applicability ofdifferent probe
geometries and found that a probe that is directed upstream parallel to the
streamlines of the undisturbed flow gives the most accurate temperature data
of the melt.
Gerrard et al. [67] pumped a Newtonian fluid (oil) through a narrow
capillary ( I , = 0.425 mm and 0.208 mm, 33 I l/r, I 459). They measured
the pressure drop, the flow rate, the inlet temperature, the wall temperatures,
and the radial temperature distribution at the exit. The calculated values of
the pressure drop and the temperature at the exit reportedly agree with the
measured values within 5%. The viscosity was taken to be a function of tem-
perature; expansion cooling was neglected in the analysis.
Mennig [72] extruded polymer melt (low density polyethylene) through a
capillary ( I , = 3.5 mm, l/r, = 225.7) at adiabatic wall conditions. Measured
quantities were the temperature in the center of the entering polymer stream,
the wall temperature distribution, the radial temperature distribution at the
VISCOUS DISSIPATION MOLTENPOLYMERS
IN FLOWING 249

exit, and the total pressure drop. The calculated values of the wall tempera-
tures and of the radial temperature distribution exceeded the measured
values by about 5%. The viscosity has been taken to be a function of shear
rate and of temperature; expansion cooling has been included in the analysis.
For capillary flow, Daryanani et al. [75] measured the average heat flux
through the wall using an electrical compensation method. From the total
pressure drop and the heat flux through the wall, they calculated the average
temperature increase between entrance and exit of the capillary.
Winter [91] extruded a polymer melt (low density polyethylene) through an
annulus ( K = 0.955 and 0.972) with rotating inner cylinder. The measured
quantities were the mass flow rate, the pressure distribution, the rotational
speed of the inner cylinder, the radial temperature distribution at the entrance
and at the exit, four temperatures each at the inner and at the outer wall. As
shown in Fig. 19 the developing temperatures have been calculated beginning
with the measured temperature distribution at the inlet. For the exit tem-
perature distribution, measured and calculated values agreed up to Y zz 0.75
within 5% of the temperature increase (at the outer wall, 0.75 IY 5 1 the
temperature distribution has not been desribed sufficiently with only four
temperature readings). The measured and calculated pressure gradients
agree within 8%. Expansion cooling has been neglected in the analysis.

FIG.19. Comparison of measured and calculated temperature profiles in helical flow [91].
250 HORSTH. WINTER

C. SHEARFLOWWITH CLOSEDSTREAM
LINES

The shear flow geometries with closed stream lines studied most widely
are circular Couette flow and its limiting case, i.e., plane Couette flow
(IC+ 1). The fluid is sheared in the annular gap between two concentric
cylinders in relative rotation to each other (Fig. 20). The axial velocity
component u, is zero. At time t = 0, the Couette system is started from rest
at isothermal conditions with a step in shear rate (Q(t I 0) = 0 and Q(0< t ) =
fo = const); alternatively the system might be started with a step in shear
stress.
Three types of development are superimposed on each other, each of them
on a different time scale:

Kinematic development: The fluid has to be accelerated until it reaches a


velocity and a shear rate independent of time. The kinetic development can
be calculated for a Newtonian fluid; a practically constant velocity field is
reached after [1111

t = ph2/16q (2.64)
(h is the gap width, q the constant Newtonian viscosity, and p the density).
For viscoelastic liquids an estimate on the duration of kinematic develop-

FIG.20. Flow geometry of circular Couette flow.


VISCOUSDISSIPATION MOLTENPOLYMERS
IN FLOWING 25 1

ment can be made from the loss and the storage modu1esG”andG’measured
in periodic shear experiments at frequency w = l / t [1121
t >> ~ ( P / G ” ) ” ~ or t >> h(~/G’)l’~ (2.65)

For startup experiments on polymer melts, the kinematic development


generally is assumed to be completed before the rheological and the thermal
development has actually started.
Rheological development: The viscosity ~ ( 9T, , t) needs some time of de-
formation at constant shear rate, until it adopts a constant value.
Thermal development: Due to viscous dissipation beginning at time t = 0,
the temperatures in the gap rise until the temperature gradients toward the
walls are large enough to conduct away all the newly dissipated energy.
Convection does not influencethe temperature field because the temperatures
along stream lines are constant.

1. Assumptions and System of Equations


The assumptions corresponding tothe ones listed in Section II.B.l are:
incompressiblefluid with constant thermal conductivity and diffusivity ;
no change in z direction;
rotational symmetry (a/a@ = 0);
velocity # 0;U, = v, = 0;
no slip at walls;
inertia negligible; kinematically developed velocity at t = 0;
gravity negligible;
viscosity measured at constant temperatures and constant shear rate gives
applicable instantaneous values of the viscosity during temperature changes
and during small changes in shear rate; rheologically developed stress at
t = 0.

The stress‘equation of motion and the energy equation become

(2.66)

(2.67)

The reference values are chosen to be the same as in the helical flow analysis:
- -
v = veVi/2; h = ra - ri; 9 = ij/h; i j = ~ 6T o,) .
252 HORSTH . WINTER

The dimensionless variables are


velocity v, = v,P,
radial position R = r/ra = (1 - K ) r / h , K IR I 1,
r - ri
y=- , 05Y11,
r, - ri
h
shear stress PR8 = Gc3==3
uvl
temperature 9 = P(T - TO).
The dimensionless form of the system of equation reads

(2.68)

pc
-- as
pE aFo
= (1 - K ) ~ [ & (asR +~ ~ )
k a
Na;(Rsxr]. v, (2.69)

The initial conditions are


w,0) = 0, V,(K 0) = V@.O(R) (2.70)
where V',JR) is the kinematically developed velocity at the initial tempera-
ture. The boundary conditions are
d 9 ( ~ Fo)
dR
,

a q i , FO)
= Bii
ss,i- ~ ( I cFo)
1--K
,
+-1 Ci- K d 9dFo
9,, - 9(1, Fo) -- C, d9(1, Fo)
,
1
( ~Fo)

= Bi, 0 = Fo. (2.71)


aR 1-u 1 - K dFo
V@(K,
Fo) = 2
Ve(1, Fo) = 0
The thermal boundary condition is an energy balance of the inner and of the
outer wall. The heat flux into the wall is equal to the heat flux out of the wall
minus the change of energy stored in the wall. The boundary condition has
already been described in Section 1I.A. It is repeated here to show the com-
plete mathematical problem at once (Bi, < 0; Bi,, Ci, C, > 0).
2. Dimensionless Parameters
For a description of most of the dimensionless parameters, the reader
is referred to Section II.B.2. The Nahme number, Eq. (2.55), compares the
dissipation term and the conduction term of the equation of energy. The
IN FLOWING
VISCOUSDISSIPATION MOLTENPOLYMERS 253

ratio of radii K shows the influence of curvature, and m describes the shear
thinning effect of the viscosity.
Instead of the Graetz number one introduces as a dimensionless variable
the Fourier number
FO = ta/h2 (2.72)
which can be understood as the ratio of the current time of the experiment
and the time needed for heat conduction from the center of the channel to the
wall. At Fo = 1-4, depending on the thermal boundary conditions, the
thermal development is completed. The Fourier number corresponds to Z in
the heat transfer problem with open stream lines, where one might define
an average residence time f = z/O,:
Z za ai
Z = = -= -
h2 = Fo. (2.73)
1 h2a,
3. Solution Procedure and Calculated Results
The 0 componentLof the equation of motion is the same for annular shear
flow with open and with closed stream lines. The kinematically developed
velocity V,( Y, 0) at isothermal conditions can be calculated with the existing
numerical program of Section 1I.B without any changes. The same is true
for the thermally developed case at large times at constant thermal boundary
conditions since the conduction and the convection terms are identical in
both types of flow. If one replaces Z by Fo and sets V,(Y) = pz =
(which is an arbitrary small value to avoid singularities in the program),
even the developing velocity V@(Y,Fo), temperature S(Y, Fo), and shear
stress PRe( Y , Fo) can formally be taken from the existing program without
further considerations; see Table VI. The capacitance parameter, however,
has to be included in the thermal boundary condition.
The solution procedure is basically the same for steady shear flow with
open stream lines and for unsteady shear flow with closed stream lines
(Couette system), and it would have been possible to treat it in one special
section in the beginning. For two reasons, however, this has not been done
in this study: (1) shear flow with open stream lines is much more important
for polymer processing; (2) the frequent change from Z to Fo would make
the explanations difficult to comprehend. The solution procedure in Section
1I.B is meant to be an example, and it will not be described repeatedly for the
corresponding problem in this section.
The geometry of a cone-and-plate or a plate-and-plate viscometer cannot
be described by the existing shear flow program. Turian and Bird [52-541
estimated the temperature effects in cone-and-plate systems by applying the
maximal gap width (at the outer radius) to a plane Couette system with
254 HORSTH. WINTER

isothermal walls. The radial heat conduction, which might diminish the
effect of dissipation, is neglected.
The development of the temperatures in circular Couette flow is a function
of the dimensionless parameters Na,Fo, IC,m,and of the thermal boundary
conditions. In Figs. 21 and 22, the influence of the geometry on the develop-

plane slit \ annulus


- 0.8 - I Y = 0.999) with x = 0.5
9
>.
e \< \
- .
-
>. - \
6
* . \
0 . A\\ \ \\
\
\ \

FIG.21. Comparison of developing temperatures for plane and for circular Couette flow.
The outer wall is taken to be isothermal; the inner wall is close to isothermal (Bii = - 100)
and close to adiabatical (Bi, = - 1). Na = 1; rn = 2; C, = 0.

I4
>
9
l%-1
2
e
c

2
E
(u 0.5
ff
6
W
c
z
-
0
P O
10-~ lo-* lo-' 1
dimensionless time Fo

FIG.22. Development of the average temperature in plane and in circular Couette flow. The
solid lines correspond to the development with both walls close to isothermal (Bi, = - 1001,
and for the dashed lines the inner wall has been taken to be close to adiabatical (Bii = - 1).
Na = 1 ; m = 2;C, = 0.
VISCOUS
DISSIPATION MOLTENPOLYMERS 255
IN FLOWING

ing temperature 9(Y, Fo) will be demonstrated for plane Couette flow
(K % 1) and for circular Couette flow with K = 0.5, both with constant
temperature of the surroundings equal to the initial temperature (9(Y, 0) =
gS,,= 9,,, = 0); the outer wall is taken to be isothermal (Bi, = co) and the
inner wall is taken to be close to isothermal (Bi, = -100) and close to
adiabatic (Bi, = -l), respectively. The thermal capacitance of the wall is
neglected (Ci = 0). For the plane slit the shear rate and hence the viscous
dissipation are nearly uniform. The temperatures rise uniformly until the
conduction toward the walls takes more and more heat out of the channel.
When the temperature gradients are large enough to conduct away all the
newly dissipated energy, the fully developed temperature field is reach. If the
inner wall is nearly adiabatical (Bi, = - l), the temperature gradient has to
adopt larger values since nearly all the dissipated energy has to be conducted
to the other wall on the outside. The corresponding temperatures for circular
Couettepow ( K = 0.5) are asymmetrical through the geometry of the system,
additionally to the asymmetry of the thermal boundary condition. The shear
rate and the viscous dissipation is much larger at the inner wall than at the
outer one. The comparison of the average temperature ~ ( F oin) Fig. 22
shows that the development is much faster if both walls are cooled instead
of one wall being nearly adiabatical (Bi, = - 1).
The thermal development depends on the capacitance of the walls. In an
example (Fig. 23) the outer wall of a circular Couette system is taken to be
isothermal (9(1, Fo) = 0); the boundary condition at the inner wall is de-
scribed by Bi, = - 1, ss,,= 0, and different values of the capacitance param-
eter Ci. The thermal development is delayed more, the larger the capacitance
of the wall is taken to be.

c , =o

>,=O , Bia=o
C,=IO

--
-

lo-’ 1 10
Fourier number Fo

FIG.23. Thermal development of circular Couette flow depending on the capacitance


parameter Ci at the inner wall; the outer wall is taken to be at constant temperature. Na = 1 ;
m = 2; &.i = 0; S,,a = 0; Bii = - 1.
256 HORSTH. WINTER

FIG.24. Developing temperature in circular Couette flow for Ci = 0 and Ci = 0.1. The
temperature of the outer wall is taken to be constant and equal to the initial temperature.
K = 0.S;Na = 1 ; m = 2;Bii = -1.

The development of the temperature near the wall is determined by the


value of C. For the example in Fig. 24 with C i = 0.1, the guide point initially
is very close to the boundary. As the inner wall heats up, the guide point
moves away from the boundary until the temperature of both the fluid and
the wall reaches its full development. Dissipation and conduction balance
and the temperature gradient at the inner wall becomes independent of Ci.
a. Unsteady Plane Shear Flow with Closed Stream Lines. Analytical studies
that include the time dependence of the viscosity q(fo, T , t ) do not seem to
be available. Several authors calculated the developing temperature field in
plane Couette flow of fluids with a viscosity independent of time:
Gruntfest [89] : q ( T )= q(To)e-B'T-To),

Krekel[86]: ~ ( j T)
, = _sinh-' -
AY ' (C;TJ7
Powell and Middleman [92] : v] = const.,

Winter [SS]:

Practical applications of their studies are the Couette rheometer [88, 89,921
and a shearing device for breaking up particles suspended in a fluid [86].
VISCOUSDISSIPATION MOLTENPOLYMERS
IN FLOWING 257

For the following example the assumption about the fully developed
stress at t = 0 will be lifted. A Couette system is kept at rest, and the stress
in the system is zero. At time t = 0 a shear experiment with constant shear
rate is started. The shear stress 7 , z ( t )is found experimentally (see for instance
[113]) to be governed by a time-dependent viscosity that increases gradually,
goes through a maximum, and approaches a constant value. If these viscosity
data are available, they might be used in the numerical program. For de-
monstrational purposes, the viscosity curve is approximated by
q($, T, t ) = [ ~ ~ ~ / ~ ) ( l ' m ) ~ ' e - " ' T --T e-'l')(l
L ,
o)](l + cZe-'l'), (2.74)
Y

Eq. (2.47)
which qualitatively fits the measured curve shapes. The maximum viscosity
is chosen to be three times the viscosity of steady shear flow; the time of the
maximum is chosen to be Fo = 0.1, i.e., at about one-tenth of the thermal
development time.
The time-dependent viscosity contains an elastic contribution, which,
however, is not specified unless one uses a complete rheological constitutive
equation. In the calculation of the dissipated energy, the elastic part of the
work of the stress is taken to be negligible compared to the viscous part.
The stress growth curve as chosen in Eq. (2.74) is reproduced by the nu-
merical program with Na = 0.001 (dashed lines in Fig. 25). If viscous dissipa-
tion is important (Na = 1, for instance) the stress reaches an earlier maximum
at a lower value; the general shape of the curve is not changed through the
I .

'\ \NO =0.001


.oN' 1

dimensionless time Fo I Fourier number 1

FIG.25. Thermal influence calculated for the startup experiment of plane Couette flow with
time dependent viscosity as described in Eq. (2.74). The walls are taken to be at constant tem-
perature equal to the initial temperature; rn = 2.5.
258 HORSTH. WINTER

effect of dissipation, and rheological and thermal effects seem undistinguish-


able in stress growth experiments. For comparison, the developing shear
stress curves for (rheologically) time-independent viscosity (as described in
Eq. (2.47)) are calculated and drawn as solid lines in Fig. 25.
b. Fully Developed Temperature Field. The fully developed case has drawn
much attention (see Table V), which is due to a double-valued solution, found
in 1940 by Nahme [44]for plane Couette flow of Newtonian fluids.
The shear stress in fully developed circular Couette flow (including plane
Couette flow as a limiting case) cannot exceed a certain value, even if the
shear rate is very large; for shear stresses below the maximum possible value,
there are always two feasible shear rates 9, a small one at high viscosity and
low temperature and a large one at low viscosity and high temperatures.
Changes from one shear rate to the corresponding one require large tempera-
ture changes, and due to the heat capacity of the system together with the
small thermal conductivity of the polymer, oscillations between the two
states do not seem possible.
For demonstrating the double-valued solution, Nahme [44] used a dimen-
sionless shear stress o* and a dimensionless shear rate $*, whose definition
can be extended to power law fluids:
t* = RdR,
Nal/(I+m)p oo)/PR@(R, O), (2.75)
j,* = Nam/(l+m) (2.76)
PRe(R, 00) and PR,(R, 0) are the dimensionless shear stress (see Eq. (2.45))
of the fully developed temperature field and of the isothermal case, respec-
tively; the ratio of the two is independent of R. The dimensionlessshear stress

dim1 schear stress T*

FIG.26. Shear rate f* as a function of shear stress 7* (both defined in Eqs. (2.76) and (2.75))
of the fully developed temperature field; the parameter is the geometry.
Viscous DISSIPATION
IN FLOWING POLYMERS
MOLTEN 259

x =0939
x =05
4..

go5
0)
c

rn
0)

e
>

1
diml. shear stress r*

FIG.27. Average temperature gm of the fully developed temperature field for different
geometries of circular Couette flow. Both walls are at constant temperature (9, = 0); m = 2.

P,,(R, co) is a monotonically descreasing function of Na, and it cannot be


used by itself to demonstrate the double-valued solution. As an example, in
Figs. 26 and 27 the double-valued solution jJ*(z*) and the corresponding
average temperature 9,(.r*) of the fully developed temperature field are
shown for circular Couette flow at K = 0.5 and IC x 1. Each shear rate p* has
only one corresponding temperature 9,.

4. Experimental Studies
The gap width of Couette systems is fairly small; and it is very difficult,
if not impossible, to measure the temperature distribution by conventional
means. The wall temperatures, however, can be measured quite accurately;
other quantities measured are the torque on the system, the rotational speed
of the cylinders, and the geometry. The double-valuedness of the shear rate
seems to have been verified by Sukanek and Laurence [55] only.
For viscosity measurements, the shear rate is prescribed and the average
velocity in plane Couette flow is taken to be ti = jh/2. The experiment should
be performed at conditions close to isothermal, which means that the Nahme
number should be as small as possible:

(2.77)

The Nahme number is proportional to the square of the gap width, i.e., the
Couette system should have a very narrow gap. Manrique and Porter [57]
built a Couette rheometer with a gap of 5 x mm; reportedly they could
eliminate the influence of viscous dissipation up to shear rates of 3 x lo6 s- '.
HORSTH. WINTER

111. Elongational Flow; Shear Flow and Elongational Flow


Superimposed (Nonviscometric Flow)

The deformation during flow can be understood as a superposition of


shear, elongation, and compression. If elongational components and density
changes are negligible, the flow is shear flow, and the corresponding heat
transfer problems can be analyzed as shown before. However, there are many
engineering applications with a flow geometry different from shear flow;
how the corresponding heat transfer problems are usually treated will be
mentioned briefly. For a more detailed description, the reader will be referred
to several examples in the literature.
Other than for shear flow, there is no accepted rheological constitutive
equation available for studying heat transfer. The proposed integral and
differential constitutive equations are mostly tested in shear experiments at
constant temperature, which might not be significant for nonviscometric flow
during temperature changes. The main reason for not applying constitutive
equations of elastic liquids is the fact that they require a detailed knowledge
of the kinematics before the stress can be determined. But for other than
Couette flow experiments, the kinematics of nonviscometric flows is not
known in advance; it has to be calculated simultaneously with the stress.
Presently a large emphasis of rheology is on solving nonviscometric flow
problems at constant temperature. Rheological analysis is not advanced
enough to incorporate temperature changes, and the present method of
solution for nonviscometric engineering problems is practically identical
with the one for steady shear flow, without care of the rheological differences.

Elongational Flow
Up to now, analytical studies on nonisothermal extensional flow have
been done by means of a temperature dependent Newtonian viscosity,
Eq. (1.13), and constant density. The studies are on melt spinning of fibers
(see, for instance, [114,115]) and on film blowing (see for instance [116,1171).
The measured stress and velocity indicate that the work of the stress a:Vv
is very small (at least for film blowing [117]), and the heat transfer seems to
be determined by convection with the moving film or thread and by con-
duction to the cooling medium.

Shear Flow and Elongational Flow Superimposed


In many different channel flows, as they occur in polymer processing, the
rate of strain contains elongational components. The fluid elements are
VISCOUS IN FLOWING
DISSIPATION MOLTENPOLYMERS 261

FIG.28. Examples of converging and diverging flow: (a) Couette flow into a converging slit,
which induces a pressure gradient for continuity reasons; (b) Couette flow in a converging
annulus; (c) Poiseuille flow into a converging pipe or a converging slit; (d) radial flow in the
gap between two parallel plates.

stretched while they are accelerated or slowed down along their paths.
Examples (Fig. 28) are Couette flow into a converging slit or annulus, flow
in a tapered tube, and radial flow between parallel plates. For describing the
stress, one commonly uses the Strokes equation, Eq. (1.13), together with
some average viscosity, or one takes the equation of the generalized New-
tonian liquid, Eq. (1.14).The results of this kind of calculation seems to give
relatively good estimates on temperature changes and viscous dissipation.
Examples are heat transfer in screw extruders (see for instance [3, 102,
118-1221), in calendering [123], during mold filling [124-1291, and in melt
solidification during flow [127-1291.
If the deviations from shear flow are small, the stress might still be defined
by the viscometric functions. An example of nearly viscometric flow is
Poiseuilleflow in a pipe with constant but irregular cross section or Poiseuille
flow in curved channels with constant cross section; the induced secondary
flow in the cross section supports heat transfer toward the walls. The secon-
dary flow, however, is very small. Whereas the improvement on the heat
transfer for polymer solutions might be up to 30% [1303, for molten polymers
(low density polyethylene in curved pipe) the influence of the secondary flow
on the heat transfer was too small to be detectable with temperature probes
in the melt [131].
Another example of nearly shear flow occurs in channels near a wall, even
if the bulk of the fluid is mainly subjected to deformations other than shear
262 HORSTH.WINTER

[132]. For steady flow, the stress at the wall is described by the three viscome-
tric functions and the wall shear rate, which of course can be determined
only from the whole flow analysis including the nonviscometric part.

IV. Summary

Heat transfer in flowing molten polymers is largely influenced by rheology,


ie., by the rheological properties of the polymer and by the flow geometry.
The rheology of steady shear flow is well understood, and hence the corre-
sponding heat transfer problems can be treated most completely. However,
heat transfer studies in flow geometries other than shear are, due to the
present lack of an appropriate constitutive equation, only possible in very
simplified form.
The most important shear flow geometries are shown to be limiting cases
of helical flow, and the corresponding heat transfer problems can be solved
with one numerical program. Two groups of heat transfer problems are
analyzed in the study: heat transfer in steady shear flow with open stream
lines (represented by helical flow with a/& = 0) and the corresponding
unsteady heat transfer problem with closed stream lines (represented by
helical flow with d/dz = 0). The problem is completely determined by six
dimensionsless parameters-the Nahme number; the Graetz number (or the
Fourier number, respectively);the ratio of the radii of the annulus; the relative
average axial velocity; the power law exponent of the viscosity; and the ratio
of length to gap width-together with the boundary conditions.
The commonly used: idealized boundary conditions are replaced by the
Biot number for describing the heat conduction to the surroundings and by
the capacity parameter for describing the thermal capacity of the wall during
temperature changes with time. The conventional definition of the Nusselt
number is not applicable to heat transfer problems with significant viscous
dissipation, and a new definition has to be introduced.
The shear dependence of the viscosity is described by a power law and the
temperature dependence by an exponential function. The temperature co-
efficient of the power law region is shown to be directly related to the activa-
tion energy of the zero viscosity.

ACKNOWLEDGMENT
The author thanks Prof. G. Schenkel for his critical advice and many helpful suggestions;
he has supported not only this work but also several specific studies of the author which were
incorporated here. The author thanks Profs. A. S. Lodge, E. R. 0.Eckert, and K. Stephan for
IN FLOWING
VISCOUS DISSIPATION MOLTENPOLYMERS 263

many critical comments and the colleagues G. Ehrmann and M. H. Wagner for helpful discus-
sions on details of the study. The Deutsche Forschungsgemeinschaft is also acknowledged for
having enabled the author to spend the time from August 1973 to November 1974 in Madison at
the Rheology Research Center which was a fruitful preparation for this work.

NOMENCLATURE

a thermal diffusivity [m2/s] S wall thickness [m]


Bi Biot number [-I, see Eqs. r time [s]
(2.21) and (2.26) T temperature [K]
Cpr C" specific heat capacity at con- T average temperature [K], see
stant pressure or at con- Eq. (2.60)
stant density [Jkg K] velocity components [m/s]
C capacitance parameter of angular velocity at inner wall
wall [-I, see Eqs. (2.26) [m/sl
and (2.30) average velocity in z direo
e internal energy [J/kg] tion [m/s]
E activation energy [J/g-mole] D reference velocity [m/s], see
Fo Fourier number, at/h2 [-I Eq. (2.38)
Gz Graetz number, &h2/al [-] dimensionless velocity com-
h r, - ri = gap width [m]; h = ponents vep, v,P, vzp
r, for circular across sec- Y coordinate in r direction, see
tion Eq. (2.32)
k thermal conductivity [J/m s Z, Z = I/(/ GI) axial coordinate
KI U pressure coefficient of viscos-
1, L l/h length of the slot ity [m2/N1, 1- '(drt/aP)r.9
m power law exponent, see Eq. B temperature coefficient of
(2.47) viscosity [K-'1, q-'(tlq/
M torque [mN] 8T)P.V
Na Nahme number, V 2 f i / k [-1, rate of strain tensor [s-!]
see Eq. (255) shear rate in simple shear
Nu Nusselt number [-I, see Eq. flow [s-'1
(2.63) s unit tensor
P pressure [N/m'], see Eq. E coefficient of thermal ex-
(1.13) pansion, - p - '(dp/dT),
P dimensionless pressure gra- W-'l
dient, see Eq. (2.43) 9 dimensionless temperature,
pRZ dimensionless shear stress B(T - T o )
components, see Eq. (2.44) azimuth coordinate
and (2.45) ratio of radii, rJra
4 specific heat flux at bound- density [kg/m3]
ary [J/m2 s] stress tensor [N/m2]
r, R = r/r, radial coordinate (note: in extra stress tensor [N/m2]
Eqs. (1.9) and (2.12), R is shear angle (see Fig. 1)
the gas law constant) first and second normal
outer and inner radius of stress function in shear
annulus [m] flow
HORSTH. WINTER
INDICES
0 initial state, reference state, i, a inner or outer boundary
or related to the zero-vis- r, R, z , Z , 0 coordinates
cosity (in a,, Po, E , ) S surroundings
02 fully developed state W wall, boundary of channel
e entrance

REFERENCES
1. G. Schenkel, “Thermodynamik, Warmeerzeugung und Warmeiibertragung in der Extru-
dertechnik,” VDI-Bildungswerk BW 2185. Ver. Deut. Ing., Diisseldorf, 1972.
2. G. Schenkel, “Kunststoff Technologie,” unpublished lecture notes, Universitat Stuttgart
(1970).
3. J . R. A. Pearson, Prog. Heat Mass Transfer 5, 73 (1972).
4. G. Schenkel, Kunstst, Gummi 7,231 and 282 (1968).
5. A. B. Metzner, Adv. Heat Transfer 2, 357 (1965).
6. J . E. Porter, Trans. Inst. Chem. Eng. 49, 1 (1971).
7. P. B. Kwant, Doctoral Thesis, Technische Hogeschool, Delft, 1971.
8. R. B. Bird, W. E. Stewart, and E. N. Lightfoot, “Transport Phenomena,” 2nd ed. Wiley,
New York, 1962.
9. J. G. Oldroyd, Proc. R . Soc. London, Ser. A 200,523 (1950).
10. J. G. Oldroyd, Proc. R . SOC.London, Ser. A 202, 345 (1950).
11. H. L. Toor, Ind. Eng. Chem. 48,922 (1956).
12. R. S . Spencer and R. D. Gilmore, J . Appl. Phys. 21, 523 (1950).
13. S . Matsuoka and B. Maxwell, J. Polym. Sci. 32, 131 (1958).
14. R. S. Spencer and R. F. Boyer, J . Appl. Phys. 17,398 (1946).
15. R. H. Shoulberg, J . Appl. Polym. Sci. 7, 1597 (1963).
16. D. Hansen and C . C. Ho, J . Polym. Sci., Part A 3,659 (1965).
17. K.Eiermann, Kunststoffe 55, 335 (1865).
18. P. Lohe, Kolloid Z . & Z . Polym. 203, 115 (1965); 204, 7 (1965);205, 1 (1965).
19. V. S . Bil and N. D. Avtokratowa, Sou. Plast. (Engl. Transl.) H10,43 (1966).
20. F. Fischer, Gummi, Asbest, + Kunstst. 7, 728 (1970).
21. W. Knappe, Adv. Polym. Sci7,477 (1971).
22. H. Wilski, Kolloid Z . & Z . Polym. 248, 861 (1971).
23. J. C. Ramsey, A. L. Fricke, and J. A. Caskey, J. Appl. Polym. Sci. 17, 1597 (1973).
24. K. Hohenemser and W. Prager, Z . Angew. Math. Mech. 12,216 (1932).
25. W. 0. Criminale, J. L. Ericksen, and G. L. Filbey, Arch. Ration. Mech. Anal. 1,410(1958).
26. A. S.Lodge, “Body Tensor Fields in Continuum Mechanics, with Applications to Polymer
Rheology,” Academic Press, New York, 1974.
27. J. Meissner, Kunststoffe 61, 516 (1971).
28. 3. Meissner, Proc. Int. Congr. Rheol., 4th, 1963 Vol. 3, p. 437 (1965).
29. V. Semjonow, Adv. Polym. Sci. 5, 387 (1968).
30. W. Ostwald, Kolloid-Z. 36, 99 (1925).
31. G. V. Vinogradov and A. Y.Malkin, J . Polym. Sci.,Part A 2,2357 (1964).
32. K. H. Hellwege, W. Knappe, F. Paul, and V. Semjonow, Rheol. Acta 6, 165 (1967).
33. L. Christmann and W. Knappe, Colloid Polym. Sci.252, 705 (1974).
34. M. D. Hersey, Physics ( N . Y . )7,403 (1936).
35. H. Hausenblas, 1ng.-Arch. 18, 151 (1950).
36. E. A. Kearsley, Trans. SOC. Rheol. 6,253 (1962).
37. D. D. Joseph, Phys. Fluids 7 , 1761 (1964).
38. B. Martin, Inr. J. Non-linear Mech. 2,285 (1967).
VISCOUS DISSIPATION
IN FLOWING
MOLTENPOLYMERS 265

39. J. C. J. Nihoul, Ann. Soc. Sci. Bruxelles, Ser. 185, 18 (1971).


40. P. C. Sukanek, Chem. Eng. Sci. 26, 1775 (1971).
41. P. C. Sukanek and R. L. Laurence, Ann. Soc. Sci. Bru-xelles, Ser. T 86, 11, 201 (1972).
42. H. Schlichting, Z. Angew. Math. Mech. 31, 78 (1951).
43. R. E. Colwell, in “Computer Programs for Plastics Engineers”(1. Klein and D. I. Marshall,
eds.), p. 183. van Nostrand-Reinhold, Princeton, New Jersey, 1968.
44. R. Nahme, 1ng.-Arch. 11, 191 (1940).
45. J. Gavis and R. L. Laurence, Ind. Eng. Chem., Fundam. 7, 232 and 525 (1968).
46. R. M. Turian, Chem. Eng. Sci. 24, 1581 (1969).
47. J. C. J. Nihoul, Phys. Fluids 13, 203 (1970).
48. P. C. Sukanek, C. A. Goldstein, and R. L. Laurence, J . Fluid Mech. 57, 651 (1973).
49. R. Kumar, J. Franklin Insr. 281, 136 (1966).
50. J. M . Wartique and J. C . J. Nihoul. Ann. Sue. Sci. Bruxelles, Ser. T, 83, 111. 361 (1969).
51. G. Palma, G . Pezzin, and L. Busulini, Rheol. Acta 6, 259 (1967).
52. R. B. Bird and R. M. Turian, Chem. Eng. Sci. 17, 331 (1955).
53. R. M. Turian and R. B. Bird, Chem. Eng. Sci. 18,689 (1963).
54. R. M. Turian, Chem. Eng. Sci. 20, 771 (1965).
55. P. C. Sukanek and R. L. Laurence, AIChE J . 20,474 (1974).
56. D. D. Joseph, Phys. Fluids 8, 2195 (1965).
57. L. Manrique and R. S . Porter, Polym. Prepr. Am. Chem. Soc.. Diu. Polym. Chem. 13 992
(1972).
58. H. Zeibig, Rheol. Acta 1, 296 (1958).
59. G . M. Bartnew and W. W. Kusnetschikowa, PIaste Kaursch. 17, 187 (1970).
60. B. Martin, “Heat Transfer Coupling Effects Between a Dissipative Fluid Flow and its
Containing Metal Boundary Conditions,” Reprint of the European Working Party o n
non-Newtonian Liquid Processing (1970).
61. H. D. Kurz, “Programm fur die Berechnung der Druck- und Schleppstromung im ebenen
Spalt.” Studienarbeit Inst. fur Kunststofftechnik, Universitat Stuttgart, 1973.
62. H. C. Brinkman, Appl. Sci. Res., Sect. A 2, 120 (1951).
63. R. B. Bird, S P E J . 11, No 7, 35 (1955).
64. H. L. Toor, Trans. Soe. Rheol. 1, 177 (1957).
65. R. E. Gee and J. B. Lyon, Ind. Eng. Chem. 49,956 (1957).
66. J. Schenk and J. van Laar, Appl. Sci. Res. Sect. A 7 , 449 (1958).
67. J. E. Gerrard, F. E. Steidler, and J. K. Appeldorn, Ind. Eng. Chem., Fundam. 4, 332
(1965); 5, 260 (1966).
68. J. E. Gerrard and W. Philippoff, Proc. Int. Congr. Rheol., 4th, 1963 Vol. 2. p. 77 (1965).
69. K. Stephan, Chem.-1ng.-Tech. 39,243 (1967).
70. R. A. Morette and C . G . Gogos, Polym. Eng. Sci. 8,272 (1968).
71. H. Schluter, Doctoral Thesis, Technische Universitat, Berlin, 1969.
72. G . Mennig, Doctoral Thesis, Universitat Stuttgart, 1969; Kunststofftechnik 9, 49, 86,
and I54 (1970).
73. N. Galili and R. Takserman-Krozer, Isr. J . Technol. 9,439 (1971).
74. G . B. Froishteter and E. L. Smorodinsky. Proc. Inr. Semin. Heat Mass Transfer Rheol.
Complex Fluids, Int. Center Heat Mass Transfer, Herzeg Novi (1970).
75. R. H. Daryanani, H. Janeschitz-Kriegl, R. van Donselaar, and J. van Dam, Rheol. Acta 12,
19 (1973).
76. G. Forrest and W. L. Wilkinson, Trans. Inst. Chem. Eng. 51, 331 (1973); 52, 10 (1974).
77. H. H. Winter, Polym. Eng. Sci. 15, 84 (1975).
78. N. Galili, R. Takserman-Krozer, and Z . Rigbi, Rheol. Acra 14, 550 and 816 (1975).
79. G . Mennig, Kunststoffe 65, 693 (1975).
80. E. M . Sparrow, J. L. Novotny, and S . H. Lin, AlChE J . 9, 797 (1963).
81. A. Seifert, Doctoral Thesis, Technische Universitat, Berlin, 1969.
82. J. Vlachopulos and C. K. J. Keung, AIChE J . 18, 1272 (1972).
266 HORSTH. WINTER

83. A. Brinkmann, Doctoral Thesis, Technische Hochschule, Braunschweig, 1966.


84. H. Rehwinkel, ”Stromungswiderstand und Warmeubergang bei nicht-Newtonschen
Flussigkeiten in Ringkanalen mit rotierendem Innenzylinder,” DFG-Abschlussbericht
No. 260/24, Deutsche Forschungsgemeinschaft , Bad Godesberg, 1970.
85. H. H. Winter, Rheol. Acta 12, I (1973); 14, 764(1975).
86. J. Krekel, Doctoral Thesis, Technische Hochschule, Karlsruhe, 1964.
87. H. H. Winter, In!. J. Heat Mass Transfer 14, 1203 (1971).
88. H. H. Winter, Rheol. Acta 11, 216 (1972).
89. I. J. Gruntfest, Trans. SOC.Rheol. 7 , 195 (1963).
90. H. W. Cox and C. W. Macosco, AIChE J. 20,785 (1974).
91. H. H. Winter, Doctoral Thesis, Universitat Stuttgart, 1973.
92. R. L. Powell and S. Middleman, In!. J. Eng. Sci. 6,49 (1968).
93. R. G. Griskey and I. A. Wiehe, AIChE J. 12,308 (1966).
94. T. H. Forsyth and N. F. Murphy, Polym. Eng. Sci. 9, 22 (1969).
95. R. G. Griskey, M. H. Choi, and N. Siskovic, Polym. Ettg; Sci. 287 (1973).
96. J . L. Ericksen, Q.J. Appl. Math. 14, 318 (1956).
97. “VDI-Warmeatlas,” 2nd ed., Ver. Deut. Ing., Dusseldorf, 1974.
98. E. R. G. Eckert and R. M. Drake, “Analysis of Heat and Mass Transfer.” McGraw-Hill,
New York, 1972.
99. J . L. den Otter, Rheol. Acta 14, 329 (1975).
100. E. Uhland, Rheol. Acta 15.30 (1976).
101. L. Schiller, Z. Angew. Math. Mech. 2, 96 (1922).
102. R. M. Griffith, h d . Eng. Chem., Fundam. 1, 180 (1962).
103. L. Graetz, Ann. Phys. Chem. 18,79 (1889).
104. A. D. Gosman, W. M. Pun, A. K. Runchal, D. B. Spalding, and M. Wolfshtein, “Heat
and Mass Transfer in Recirculating Flows,” Academic Press, New York, 1969.
105. A. G. Fredrickson and R. B. Bird, Ind. Eng. Chem. 50, 347 (1958).
106. W. Nusselt, Z. Ver. Dsch. Ing. 54, 1154 (1910).
107. W. Tychesen and W. Georgi, SPE J . 18, 1509 (1962).
108. H. Janeschitz-Kriegl, J. Schijf, and J . A. M. Telgenkamp, J . Sci. Instrum. 40,415 (1963).
109. G. Schenkel, DOS 1,554,931 (1966).
110. J. van Leeuwen, Polym. Eng. Sci. 7 , 98 (1967).
1 1 I . H. Schlichting, “Boundary Layer Theory,” p. 65. McGraw-Hill, New York, 1955.
112. J. D. Ferry, “Viscoelastic Properties of Polymers,” 2nd ed., p. 121. Wiley, New York,
1969.
113. J. Meissner, Rheol. Acta 14, 201 (1975).
114. Y.T. Shah and J. R. A. Pearson, Ind. Eng. Chem., Fundam. 11, 145 (1972).
115. S. Kase, J. Appl. Polym. Sci. 18, 3267 (1974).
116. C. J. S. Petrie, AZChE J. 21, 275 (1975).
117. M. H. Wagner, Rheol. Acta 15.40 (1976).
I 18. B. Martin, J. R. A. Pearson, and B. Yates, Vniv. Cambridge, Polym. Process. Res. Cent.
Rep. No. 5 (1969).
119. R. T. Fenner, “Extruder Screw Design,” Iliffe, London, 1970.
120. 2. Tadmor and 1. Klein, “Engineering Principles of Plasticating Extrusion,” Van Nos-
trand-Reinhold, Princeton, New Jersey, 1970.
121. G. Schenkel, Kunsfstofftechnik 12, 171 and 203 (1973).
122. R. V. Torner, “Grundprozesse der Verarbeitung von Polymeren,” VEB Dtsch. Verlag
Grundstoffind. Leipzig, 1974.
123. V. J. Petrusanskij and A. I. Sachaev, Uch. Zap. Yarosl. Tekhnol. Inst. 23 (1971); cited
by Torner [ 122).
VISCOUSDISSIPATION
IN FLOWING
MOLTENPOLYMERS 267

124. J. L. Berger and C . G. Gogos, Polym. Eng. Sci. 13, 102 (1973).
125. M. R. Karnal and S. Kenig, Antec 18, 619 (1972).
126. H. H. Winter, Polym. Eng. Sci. 15,460 (1975).
127. E. Broyer, C. Gutfinger, and Z. Tadmor, Trans. SOC. Rheol. 19,423 (1975).
128. J. Rothe, Doctoral Thesis, Universitat Stuttgart, 1972
129. C. Gutfinger, E. Broyer, and Z. Tadmor, Polym. Eng. Sci.15, 515 (1975).
130. D. R. Oliver, Trans. Inst. Chem. Eng. 47T, 8 (1969).
131. H. H. Winter, unpublished experiments.
132. B. Caswell, Arch. Ration. Mech. Anal. 26. 385 (1967).
133. H. Rehwinkel, Doctoral Thesis, Technische Universitlt Berlin (1970).

You might also like