Vous êtes sur la page 1sur 33

Journal of Non-Newtonian

Fluid Mechanics,

Elsevier

B.V., Amsterdam

Science Publishers

49 (1993) 141-113

141

A novel hybrid numerical technique to model 3-D


fountain flow in injection molding processes
B. Friedrichs
Department
(Received

and S.I. Giigeri

of Mechanical
October

Engineering,

University of Delaware,

8. 1992; in revised form February

Newark,

DE 19716 (USA)

15, 1993)

Abstract
A novel hybrid two-dimensional
(2-D)/three-dimensional
(3-D) numerical technique is presented to model 3-D fountain flows in thin cavities as
encountered
in injection molding processes. At the fountain flow region,
where all three velocity components
are significant, the governing 3-D fluid
flow equations are solved by using a pressure Poisson formulation.
Behind
the flow front, where out of plane flows are negligible, the 2-D Hele-Shaw
formulation
is employed,
largely reducing the number of unknowns
in
comparison
to a fully-three-dimensional
formulation.
Boundary fitted coordinate systems (BFCS) together with the finite difference method (FDM)
are used to solve the governing equations
on a non-staggered
grid. The
formulation
is capable of handling non-linearities
in the material behavior
due to the shear-thinning
characteristics
of typical resin systems. Results are
presented
for isothermal
flow of Newtonian
and shear-thinning
fluids
through diverging and converging flow sections.
Keywords: fountain flow; grid generation;
molding; three-dimensional
analysis

Hele-Shaw

flow; hybrid 2-D/3-D

method;

injection

1. Introduction
Modeling of injection molding has been an area of considerable
research
over the last two decades. These studies resulted in several computational
algorithms that assist in the design and fabrication
of plastic parts using
injection molding processes [ 11. Modeling of the flow in these processes
represents several major challenges since the flow is inherently transient,
* Corresponding
author. Present address: Mechanical Engineering,
Chicago, Chicago, Illinois 60607, USA.
This paper is dedicated to Professor Hermann Janeschitz-Kriegl
0377-0257/93/$06.00

1993 - Elsevier

Science Publishers

University

of Illinois at

on his 70th birthday.

B.V. All rights reserved

141

B. Friedrichs and S.I. Giigeri 1 J. Non-Newtonian

Fluid Me&.

49 (1993) 141-l 73

non-isothermal,
non-Newtonian,
and includes
a free surface
moving
through cavities of highly irregular geometries.
Due to limited computational resources, the three-dimensional
(3-D) flow problem has customarily
been simplified to a two-dimensional
(2-D) problem, based on the observations of Hele-Shaw [2]. In this approach the cavity is assumed to be thin,
and out of plane flows are neglected. It was first proposed
for flow of
polymer melts by Richardson [3] who looked at isothermal, Newtonian flow
in simple geometric shapes. Over the years the Hele-Shaw formulation
has
become the basis for many computational
codes that either use a Eulerian
mesh together with a control volume (CV) approach (also referred to as
flow analysis network (FAN)) to track the motion of the flow front [4- lo]
or remesh either locally at the flow front [ 1 I] or completely [ 12- 141 such
that the mesh coincides at all times with the domain covered by the fluid.
The Hele-Shaw formulation,
however, cannot capture the details of the
flow field whenever out of plane flows are present such as in the vicinity of
the flow front (see Fig. 1). Here, fluid typically spills outward towards the
walls due to the no-slip situation
leading to the fountain
flow phenomenon first called as such by Rose [ 151 and experimentally
observed in
injection molding as reported in Refs. 16-18.
Whenever the transient location of fluid particles is of importance,
as for
temperature,
degree of cure, or short fiber orientation
predictions,
the
incorporation
of the fountain
flow phenomenon
becomes necessary to
achieve satisfactory accuracy in process modeling, see e.g. Ref. 19. Although
the fountain flow can be partially accounted
for in Hele-Shaw flows by

Hele-Shaw
(2-D flow)

Fig. 1. Flow in a thin cavity.

fountain flow
(3-D flow)

B. Friedrichs and S.I. GiiGeri 1 J. Non-Newtonian

Fluid Mech. 49 (1993) 141 -I 73

143

following heuristic approaches


such as that of Lord and Williams [20],
much attention has been directed towards a more accurate two-dimensional
analysis of the flow in the gapwise direction only. Early contributors
to the
field were Bhattarcharji
and Savic [21] who developed an analytical model
for isothermal, Newtonian flow in a pipe that is dragged over a stationary
piston. Further analytical work that relates to injection molding processes
was done by Tadmor [22] for an isothermal power-law fluid and by Castro
and Macosko [ 191 and Manas-Zloczower
et al. [23] for a reactive fluid. It
should be noted that all these studies assume a flat flow front.
The shape of the flow front was allowed to develop freely in simulations
where a tube or parallel plates are sliding over the fluid with a stationary
flow front. While Givler et al. [24] and Coyle et al. [25] solved this problem
for an isothermal, Newtonian fluid, Mavridis et al. [26] carried the modeling
to isothermal power-law fluids. The same authors later extended the work to
non-isothermal
Leonov fluids [ 271.
The advance of a fluid into a stationary cavity represents a more complex
problem
since the domain is continuously
deforming
and enlarging. A
remeshing procedure was first employed by Behrens et al. [28] who solved
for the case of an isothermal, Newtonian fluid. A similar technique was also
used by Mavridis et al. [29] who solved for two Newtonian fluids forming
a weld line and by Fauchon
et al. [30] for a Newtonian
fluid moving
through a corner. More recently Hayes et al. [31] solved for the advancement of a reacting fluid into the two-dimensional
cross-section
of a T-joint.
A stationary
mesh in conjunction
with the marker-and-cell
(MAC) technique was used by Gogos et al. [ 181 to model the advancement
of a
power-law fluid. Kamal and co-workers [32, 331 utilized the same approach
for non-isothermal
flow of a White-Metzner
fluid. While some of these
studies treat the problem in detail, it should be noted that this kind of
analysis does not give any information
about the nature of the planar flow
through the mold cavity. Therefore, the predictions can only give qualitative
understanding
in a true three-dimensional
flow situation.
It is the purpose of this paper to bridge the gap between existing in-plane and
gapwise analyses by solving for the fully three-dimensional
flow field. To
achieve computational
economy a novel hybrid numerical technique is proposed that only solves for the 3-D flow field wherever all three-velocity
components are significant; elsewhere, the 2-D Hele-Shaw formulation is retained.
2. Governing equations
2.1. 3-D Jlow region
Typical Reynolds numbers encountered
in injection molding processes
using polymers are much less than unity [ 341. Therefore,
the inertia terms

144

B. Friedrichs and S.I. Giigeri 1 J. Non-Newtonian

Fluid Mech. 49 (1993) 141-I 73

can be neglected, leading to a quasi-steady-state


creeping flow type formulation. In a three-dimensional
flow configuration
the governing equations
become
continuity:

V *24= 0,

(1)

momentum:

V*o=O;a=z-pl

(2)

r = r(j)?.

(3)

rheological

behavior:

The fluid behavior is modeled as generalized Newtonian,


i.e. z,, = 2q(j)&/
ax, z, = y(lj)(au/ay + au/ax), . . . . Amongst various possibilities, the highly
shear-dependent
viscosity can be represented by the Carreau model [35]
yI =

qcc,+ (qo- &)[ 1 + (a#]-

)2),

where y1 is the dimensionless


power-law index which is less than unity for
shear-thinning
fluids. It ranges from 0.2 to 0.8 for most polymer melts.
Furthermore,
i is the square root of i times the second invariant of the rate
of strain tensor, I a time constant
and q. and q= are the Newtonian
viscosities at zero and very high shear rates, respectively. For the Newtonian
case y1 is equal to unity.
It should be noted that the governing
equations
include no explicit
expression for the pressure which acts as a source term in the momentum
equation and adjusts itself in a way to satisfy continuity everywhere in the
domain [36]. In order to overcome
the difficulties associated
with the
missing explicit equation for pressure, the pressure Poisson equation formulation of Harlow and Welch [37] has been adopted. The first step in this
approach is to consider the transient momentum
equation

au

pz+vp

Then,

v*

=v-z.
taking

the divergence,

p;+vp
=v.z,

changing the order of differentiation


this equation becomes
0
f

and considering

a finite time increment,

v.~n+_v.Un

v*p =V*(V*z)

-p

At

(7)

To ensure continuity
at the new time step 12+ 1, the first term in the
numerator
on the right-hand
side of eqn. (7) is set to zero. Equation (7)
now gives the expression needed for pressure.

B. Fr~~drichs and XI. G&eri 1 J. ~o~-~e~~onian

Fluid Mech. 49 (1993) 141 I 173

14.5

This method is suitable for transient [37] as well as for steady-state


[38-401 problems. It should be noted that for (quasi) steady-state problems,
the transient term in the momentum equation (5) is only reintroduced as
part of the pressure Poisson fo~ulation
and does not alter the (quasr)
steady-state character of the original problem. Equation (5) is then marched
in time while updating the pressure (eqn. (7)) and the viscosity (eqn. (4))
until (quasi) steady-state is reached. The transient of this marching procedure, however, have no physical meaning. Once (quasi) steady-state is
reached, the flow front is advanced and eqns. (S), (7) and (4) are solved
again.
It should be noted that the pressure Poisson equation (7) is not obtained
from physical principles but derived mathematically;
therefore, the
boundary conditions on the pressure need to be consistent with the original
governing equations which only allow for the specification of either the
velocity or the traction on the boundaries [41]. These boundary conditions
can be obtained by taking the normal component of the momentum
equation as
Vj?*n=(V*r)*n,

(8)

leading to Neumann type conditions


straints to the system [42].

which introduce

no additional

con-

region

2.2. 2-Djow

Whenever the gapwidth is small in comparison with the planar dimensions, and the gapwise velocity component is negligible, an order of magnitude analysis shows that the momentum equations (2) can be reduced to

-.&-+& [,,,,(,,,,,lco,

_ aP(x> Y)

x momentum:

(9)
aP(x9 Y)
-~+~[ai;(x~~,-)]=O,

y momentum:

leading to the Hele-Shaw formulation.


Integrating eqn. (9) with respect to z twice and using the no-slip
boundary condition at the top and bottom walls, i.e. u = v = 0 at z = &h/2,
gives
hi2

8P
4x,

Y,

z)

ax

aP
D(X,p, 2) = - -

ay

rw*

d,-*

sr

3
Y

s
m

&

--y

B. Friedrichs and XI. GiiGeri 1 J. Non-Newtonian Fluid Mech. 49 (1993) 141-I 73

146

Defining

fluidity

as

h/2$ dZ
-

S(x, J) =
s0

YI

the mean velocities

(11)

can then be found

by performing

integration

2s ap
U=-Tax7

(12)

2s ap
=-we

Defining

the stream

function,

which satisfies continuity,

w
u=Y

V=-&,

as

(13)

and combining
stream function

it with eqn. (12), a unified


is obtained as

flow equation

in terms of the

(14)
Although this equation can also be written in terms of the pressure, the
stream function formulation
is preferred in this study, because a prescribed
flow rate boundary condition is applied at the inlet.
3. The hybrid solution technique
Experimental
observations
have shown that the fountain
flow region
extends only of the order of one gapwidth backward from the free surface
[25]. This region is, therefore,
small in comparison
to the planar length
scales of the flow field (see Fig. 1). Thus, it is advantageous
to only solve for
three-dimensional
flow in the fountain flow region using a three-dimensional
mesh which is trailed by a 2-D mesh to solve for the Hele-Shaw flow. Figure
2 depicts a typical hybrid mesh generated for a planar how. This hammerhead approach
therefore
greatly reduces the number of nodes in the
system, thus rendering a 3-D fountain flow analysis suitable for workstation
based computations.
A critical decision in constructing the hammerhead
mesh is the determination of the pseudo-boundaries
separating the two- and three-dimensional
zones. Based on the experimental
observations
and the numerous numerical
studies on 2-D fountain flow mentioned before, a length of 1.5-3 times the

B. Friedrichs and S.I. Giiqeri 1 J. Non-Newtonran

Fluid Mech. 49 (1993) 141-l 73

147

Flow

-(Hele-Shaw

2-D Mesh
approach)-,
3-D Mesh
(primitive variable approach)

Fig. 2. Hammerhead
situation.

mesh resulting

from the hybrid

solution

technique

in a fountain

gapwidth was chosen for the extent of the 3-D flow region backward
the contact line. The results later confirm this choice.

flow

from

4. Boundary conditions
Figure 3 depicts various boundary conditions that need to be specified to
solve the governing eqns. (5) and (14) for the fountain flow problem. At the
inlet, the volume flow rate is specified leading to a Dirichlet type boundary
condition for the stream function $. The boundary conditions at the side
walls in the Hele-Shaw domain are also of Dirichlet type leading to full slip.
This is consistent
with the simplifications
introduced
in the momentum
equations (9) which imply that the flow is unbounded
in its plane. At the

interface: u = u(x,y,z)
v = v(x,y,z)
W=Q
I

interface: g=

Cl(Y)

Top and bottom wall:


u=v=w=o

side walls (Hele-Shaw):


w = const.

Fig. 3. Boundary

conditions.

side walls (3-D): v_- Q= 0


(g._n)*t,=O
(a- _n).t,=o

B. Friedrich

148

and S.I. Giiqeri 1 J. Non-Newtonian Fluid Mech. 49 (1993) 141~ I73

top and bottom walls, the no-slip condition holds which was used for the
integration of the momentum equations (9). This interface between the 2-D
Hele-Shaw and the 3-D flow domain behind the flow front is formed by an
overlap of one grid layer in the direction of the flow. At the downstream
end of the overlap, the normal derivative of the stream function is computed
from the average velocities in the 3-D flow region using eqn. (13).
At the upstream end of the overlap the planar velocity components
u and
u are computed
from the stream function distribution
using eqn. (10)
written in terms of the stream function instead of pressure. Naturally,
the
gapwise velocity component
w is zero. These conditions form the upstream
boundary conditions for the 3-D flow region. At the top and bottom wall of
that region, the no-slip condition is maintained.
To be consistent with the
Hele-Shaw formulation,
a slip boundary
condition
is applied at the side
walls which provides two boundary conditions. The third condition comes
from the impermeability
of the cavity boundaries.
Since the pressure drops
out from these equations, no additional boundary condition is needed. At
the free surface a force balance gives [29,43]
(15)
where a and b denote positions on either side of the free surface (see Fig. 3).
Ca is the capillary number defined as the ratio of viscous to surface tension
forces
Ca

=gU,

(16)

For creeping flow of highly viscous fluids this number is large (values for
the surface tension y of various polymer melts can be found in Ref. 44) and
eqn. ( 15) can finally be written as
o-n

=o,

which needs to be
four equations for
the pressure at the
determined
from
previous discussion
equation.

(17)
solved together with the continuity equation to provide
the four unknowns U, 21,w and p. Although this specifies
free surface, it is worthwhile to note that the pressure is
the continuity
constraint
which is consistent
with the
on the boundary
conditions for the pressure Poisson

5. The moving contact line


Because of the no-slip boundary
condition
specified at the top and
bottom walls of the cavity, there exists an infinite force singularity at the

B. Friedrichs and S.I. Giigeri 1 J. Non-Newtonian

contact point at time 1,

Fluid Mech. 49 (1993) 141-l 73

149

contact point at time tn+,

flow

Fig. 4. Flow front

advancement.

moving contact line. For a review of this problem see Dussan [45]. According to the same author, the problem is not a kinematic but a dynamic
incompatibility,
among the modeling assumptions such as no-slip at the wall
and no stresses at the free surface. To the best of our knowledge there is no
model that draws on the physics and resolves the issue. One way to remedy
this problem is to introduce an artificial slip coefficient at the vicinity of the
contact line [27,33,46-481. In this work, however, the approach of Behrens
et al. [ 281 has been adopted which avoids the computation
of the velocities
at the contact line. Instead, the contact point is held fixed while the flow
front is advanced using the computed velocities at time t, (Fig. 4). Then, a
curve is fitted to the new location of the nodes along the flow front. The
intersection of the curve with the cavity wall forms the new contact point at
time tn+,. Thus, every position of the cavity wall is covered by a different
fluid particle which resembles a rolling motion of the fluid as suggested by
Dussan and Davis [49]. Although strict convergence of the solution cannot
be established under strong spatial refinement,
second-order
convergence
with respect to the flow rate at the entrance and the free surface was found
for the practicable range of lo-30 nodes through the gap [50]. This suggests
that the solution is sufficiently accurate with respect to the velocity field in
this range of nodes. It should also be noted that the amount of fluid which
appears to be lost through the walls computationally,
is small in comparison
with the discretization
error for the overall problem and can therefore be
neglected [28,50].
6. Numerical implementation
6.1 Grid generation
Boundary fitted coordinate systems (BFCS) in conjunction
with the finite
difference method (FDM) were chosen for the solution of the governing
equations
in which an irregularly
shaped domain in the physical x, y, z
space is mapped into a regularly shaped domain in the computational
5, r,~,v

B. Friedrichs and XI. Giigeri / J. Non-Newtonian

150

Physical

Domain

Fluid Mech. 49 (1993) 141-173

Computational

Domain

Fig. 5. 3-D mapping of an irregularly


computational
domain.

shaped

physical

domain

into

a regularly

shaped

space (see Fig. 5). Numerical grid generation as proposed by Thompson et


al. [51,52] is used to generate the grid in the physical domain. The grid
generation equations for the 3-D flow domain can be written as
v< =gP,
V2q =gQ,

(18)

V,t = g=R,
where P, Q and R are grid control functions that enable the control of grid
line concentration.
The quantities g, gz2 and g3 are the diagonal components of the contravariant
metric tensor and defined as
g = v;l

. V(,

(19)

where 5 = [, ye, v for i = 1, 2, 3, respectively.


independent
variables, eqn. ( 18) becomes
g(x;:

+ Px,) +g(x,,,

dependent

+ Pyt) +gz2(yqq +

= 0,

Q.Y,J +g33(~,,,. + Ryv)

(20)

+ 2(giZyc:, +g3_v,;,,+ gZ3y,,.) = 0,


g(z<; + P,;) +g=(+

and

+ Qx,) +g3(x ,,,,+ Rx,.)

-t 2(g %~o + g?Ycv + g%,,)


g(y,,

Interchanging

+ Qz,, +gj3(zv,, + Rz,)

+ 2(g ~~~,]+ g 3z;,,+ g~3z,zr)= 0,


which can be solved using second-order

finite difference

expressions.

B. Friedrichs and XI. Giiferi 1 J. Non-Newtonian Fluid Mech. 49 (1993) 141-l 73

151

The boundary conditions in the present study are of Dirichlet type, such
that grids need to be specified on the bounding surfaces of the domain. These
can be generated on a v = const surface, for example, by using [52]:

(21)
which are similar to eqn. ( 18) except that the Laplacian is replaced with the
second-order
Beltramian.
Interchanging
dependent
and independent
variables renders
gl(xC; + Px;) + g2(x,, + Qx,) +g12xrq = n;(k, + k,,),
+

Qv,> +g2ysq = &(k + k,h

gh

+ Pyt) +gb,,

g(+

+ Pq) + g(z,, + Qz,) + gz;, = n;(k, + k,,),

(22)

which, again, can be solved using centered, second-order


finite difference
expressions. In the above expressions n;, n; and n; are the components
of
the outward unit normal vector to a v = const. surface and k, + k,, is the sum
of the principal curvatures to be evaluated from an equation of the form
F(x, y, z) = J?. But on the free surface such a function is not known. Although
it is, in principle, possible to determine it locally from the location of
surrounding nodes, this is too inaccurate for the present purpose. Instead, the
nodes on the free surface are merely redistributed
along the two varying
computational
coordinates
on that surface according to the specified distribution along the bounding edges. This avoids undesirable node concentration
due to the motion of the flow front.
A particularly useful characteristic of the above generation equation is that
the grid control functions P, Q and R can be determined from the distribution
on the edges of the domain and then interpolated into the interior using linear
(2-D) and transfinite
(3-D) interpolation.
The control function
P, for
example, can be found in l-D, i.e. along a varying < edge, as [52]
p=_*

(23)
x;
which follows directly from eqn. (20).
Node concentration
along a varying 5 edge towards the i = m node, for
example, can be achieved by using an expression of the form [53]

$+3(q,

(24)

where s, is the total length, s the incremental length and B the concentration
factor. In this study eqn. (24) has been used to concentrate
nodes towards

152

B. Friedrichs and S.I. GiiGeri / J. Non-Newtonian Fluid Mech. 49 (1993) 141-I 73

the contact lines in the direction of the flow. Additional


information
on
three-dimensional
grid generation and mapping in materials processing can
be found in Refs. 50 and 54.
The generation
of the grid in the 2-D Hele-Shaw domain follows along
the same lines and is therefore not discussed here any further.
6.2 Transformation

and discretization

of the governing equations

In order to achieve maximum accuracy for a given number of nodes and


possible high viscosity gradients, the governing equations (5) (7) and (4)
should be solved in their conservative
form [55]. In other words, the
coefficients of the derivative terms, such as viscosity, must either be constant
or, if variable, their derivatives may not appear anywhere in the partial
differential
equation (PDE) [56]. Also to retain conservation
in the discretized version on an irregularly
shaped domain, the derivatives
of a
quantity fin 3-D, with respect to the physical coordinates x, y and z have to
be stated in their geometrically conservative form which can be written as [ 571

JJ;- = (fJL>t + (fJ~ls>~+ (fJv.A,


Jfi = (fJi,>t+ CLJry),+ (fJy&

(25)

JL = (fJSz>t+ (fJ~z>~+ (fJv,h,


where J is the Jacobian
of the transformation
and c,, q,, . . . are the
components
of the contravariant
base vectors which can be written in terms
of the covariant base vectors as
V+z,

x a,+)

J = a, - (a2 x a3).

(i = 1, 2, 3)

(i, j, k) cyclic,

(26)

Here, a = (x~,, yrl, z;,. These expressions permit the cancellation of the fluxes
across common cell boundaries, thus ensuring overall conservation.
For second-order
derivatives eqns. (25) need to be applied in a consecutive
manner. For example, ar, /ax, is discretized as
a
a.y LX

(27)

B. Friedrichs and XI. Giigeri / J. Non-Newtonian

Fluid Mech. 49 (1993) 141-l 73

153

It is important
that the metrics at the half-integer
points may not be
averaged but computed.
Otherwise, a uniform flow field is not preserved
[57]. Therefore, the nodal coordinates
of the half-integer points need to be
found by, for example, averaging. All other variables, such as viscosity, can
be computed by averaging among the values at the full integer nodes. The
discretization
of Neumann type boundary conditions follows along the same
lines. A good review on the topic can be found in Ref. 57.
6.3 Solution

technique for algebraic equations

The momentum
equations are solved using the explicit Euler scheme for
the time derivative. The maximum time step is limited by [58,59]

At<

-+
r.J,k

>min

if 4,,,k < 0,

(28)

is the multiplication
factor of the velocity component of interest
where
Al.J,k
in the discretized version of the right-hand side of eqn. (5). Although this
time step is small, it is similar to the process time and other techniques such
as fully implicit or approximate
factorization
show no savings in CPU time
[50]. The convergence criterion was chosen to be
Un++Un

un

I-.
<IO_4

The pressure Poisson equation


and the grid generation
solved using successive over-relaxation
(SOR).

(29)
equations

are

6.4 Overall solution procedure


The flow simulation is initialized with a small prespecified mesh at the
inlet and an initial guess for the flow variables. The momentum equation (5)
is then marched in time while updating the pressure using the Poisson
equation (7). During this procedure the viscosity and the stream function in
the Hele-Shaw domain are updated periodically after a specified number of
velocity updates. Once convergence is achieved the nodes on the free surface
are advanced using the first-order explicit scheme
AX = u At,
Ay = v At,

(30)

AZ = w At,
while holding the nodes along the contact line fixed, as mentioned before. It
should be noted that the time step in eqn. (30) is user-specified
and only
used to advance the flow front. The momentum
equation (5), instead, is

154

B. Friedrichs and S.I. Giiqeri / J. Non-Xewtoniart

Fluid Mech. 49 (1993) 141-l 73

marched using a different time step computed from eqn. (28). Every
coordinate line on the free surface that may intersect the cavitys top and
bottom wall is fitted using Akimas method 1601. This is a one-dimensional
interpolation from a set of given data points using piecewise third-order
polynomials which has proven to produce smooth results. The new contact
points, i.e. the new contact line, is found by intersecting the fitted curves
with the cavity walls. To avoid undesirable node concentration, the nodes
along each computational
coordinate line on the free surface are redistributed algebraically. Finally, the mesh is regenerated on the boundaries
and in the interior using the new position of the flow front. Whereas a 3-D
mesh is generated downstream of the interface of the 2-D and 3-D flow
domain, a planar 2-D mesh is generated upstream in the Hele-Shaw
domain. Using the field variables from the previous time step as the initial
guess for the next time step, the procedure is repeated until the cavity is
filled.
Since the domain is continuously deforming and enlarging, lines of nodes
are added periodically in the downchannel direction of the Hele-Shaw
domain to avoid either too fine or too coarse meshes. To achieve this, nodes
are distributed along the cavity side walls which not only specify the cavity
shape but also trigger the addition of node lines when passed by the flow
front. It should be noted that the number of grid nodes in the 3-D flow
domain is held constant due to only small changes in its size in the
downchannel direction.
7. Results
Isothermal flow of a Newtonian and a shear-thinning fluid through
diverging and converging cavity sections of constant gapwidth is considered
to demonstrate the proposed technique. Results are presented for the
shear-thinning fluid unless stated otherwise. The viscosity of the Newtonian
fluid was chosen to be 100 Pas. Material data for the Carreau model were
taken for a polystyrene (PS) melt at 18OC with q. = 14 800 Pas,
qL = 0 Pa.s, 3, = 1.04 s, IZ= 0.4 [ 121. The volume flow rate at the inlet was
chosen to be 200 cm3 s- for the diverging flow configuration in order to
obtain wall shear rates of 0 ( IO s-) found in typical injection molding
processes. Figure 6(a) shows the top view of the cavity shape and the initial
mesh at the inlet. Due to symmetry, only the left half is considered. Figure
6(b) shows an intermediate mesh while Fig. 6(c) shows the final mesh after
0.3 s of mold filling using a maximum time step of 0.001 s for the flow front
advancement. An enlarged side view of the downchannel symmetry line of
the final mesh is depicted in Fig. 6(d). Here, the interface between the 2-D
and the 3-D flow region is clearly seen. Although only a planar 2-D mesh

B. Friedrichs and S.I. Gibgeri /J. Non-Newtonian Fluid Mech. 49 (1993) 141-173

1.55

is needed in the Hele-Shaw domain, a grid is generated algebraically in the


gapwise direction. The nodes of this grid form the collocation points for the
fluidity integral given by eqn. (11) and the velocity integral given by eqn.
(10). This grid, however, is not needed for the stream function computation,

Fig. 6 (a).

Fig. 6 (b).

156

B. Friedrichs and S.I. Giiceri 1 J. Non-Newtonian Fluid Mech. 49 (1993) 141-I 73

interface
2-D

r-1

3-D

Fig. 6. (a) Initial mesh and boundary of cavity; (b) intermediate


mesh; (c) final mesh; (d)
side view of final mesh along the downchannel
symmetry line at the flow front.

thus reducing the mesh in Fig. 6(d)to the hammerhead


mesh in Fig. 2. It
should be noted that grid lines are concentrated
strongly toward the flow
front using eqn (24). In order to generate a high quality grid in the
semicircular part of the fountain flow region, the last 4 grid lines/layers in
the downchannel
direction are collapsed into the contact point/line (see Fig.
6(d). The initial mesh was chosen to have a semicircular flow front with 1.5
nodes in the gapwise direction.
Figure 7 shows the dimensionless
flow front elongation which is defined
as
Elongation

Downchannel

distance

between tip the flow front and contact


half gapwidth

line
(31)

B. Friedrichs and XI. Giiqeri 1 J. Non-Newtonian

00

0 00

0 05

Fluid Mech. 49 (1993) 141 -I 73

010

time

Fig. 7. Dimensionless

flow front elongation

0 15

020

157

025

0 30

[s]

for diverging

flow configuration.

For a flat flow front this ratio is zero while it is unity for a semicircular
front. Values for the dimensionless flow front elongation were taken at the
center of the flow front. It is seen that the front shape remains close to
semicircular
in accordance
with Hoffman [61] and experimental
observations for flow in a tube [28]. Figure 7 also shows that in the straight section
of the channel a constant value of 1.Ol for the Newtonian and 1.15 for the
shear-thinning
fluid are held rather steadily after some initial oscillations.
They only drop slightly after the diverging section has been reached. These
values correspond well to a grid refinement study performed for 2-D parallel
plate flow using the same formulation
[50]. It is seen in Fig. 8 that for a
Newtonian fluid using between 15 and 41 nodes in the gapwise direction, the
values lie well within the reported range of 0.84- 1.04 in other numerical
studies [25,26,28,29,30,62,63].
Up to now, no experimental
data have been
reported
for the elongation
in parallel plate flow. Behrens et al. [28],
however, were able to simulate the elongation to within the error of their
experiments for pipe flow. Using the same numerical technique for parallel
plate flow they predicted a value of 0.94. The results of this work are seen
to converge towards this value. In addition, no experimental data have been
reported
for shear-thinning
fluids at realistic injection molding process
conditions. But, the asymptotic value of 1.04 reached for the shear-thinning

B. Friedrichs and S.I. Giigeri / J. Non-Newtonian Fluid Mech. 49 (1993) 141-I 73

158
14

:
z

oa-

3
0
c

06-

:
al
-E

04-

0
.-

E
;

.-0

02-

00

of

I
40

I
30

I
20

10

nodes

in

Fig. 8. Convergence of flow front elongation

the

gapwise

direction

for 2-D parallel plate flow.

corresponds
well to the work of Mavridis
et al. [26] who
predicted a value of 1.O for a power-law fluid with an exponent of 0.5 under
similar flow conditions.
Figure 9(a) shows the velocity u in the .X direction at the end of the
simulation
for the shear-thinning
fluid. The drawing is enlarged in the
gapwise direction for clarity. Thus, the flow front is stretched into an almost
straight line. Only the bottom half of the domain is displayed due to
symmetry; therefore, the top surface in Fig. 9(a) corresponds
to the midplane of the cavity. Acceleration
of the fluid is noticeable
at the bend
forming the entrance to the diverging section. The parabolic velocity profile
and the slowdown of the fluid in the fountain flow region can be observed
at the side. Figure 9(b) shows this more clearly in a side view of the
fountain flow region. Here, both the top and bottom half are displayed and
the figure is enlarged in every direction so that the real flow front shape is
depicted. Figure 9(c) depicts the same situation for the Newtonian fluid. It
should be noticed that the upstream,
parabolic
velocity profile is more
protruding.
Moreover,
the deceleration
occurs closer to the flow front.
These differences will have a large impact on fluid particle relocation
as
demonstrated
later.
polystyrene

B. Friedrichs and XI. G&eri 1 J. Non-Newtonian Fluid Mech. 49 (1993) 141-l 73

159

Fig. 9. (a) Velocity component u at the bottom half of the cavity for the shear-thinning
fluid
(min. O.Om s-I max, 1.614ms~);
(b) side view of (a) at the flow front; (c) side view for
Newtonian
fluid (min, 0.0 m s-l, max, 1.423m S-).

Fig. 10. (a) Velocity component


u at the bottom
fluid (min, 0.0 m s-; max, 0.557 m SC).

half of the cavity

for the shear-thinning

160

B. Friedrichs and S.I. GiiGeri 1 J. Non-Newtonian

Fluid Mech. 49 (1993) 141-l 73

The other planar velocity component


u in the y direction is depicted in
Fig. 10. Looking at the mid-plane, it can be observed, again, how the fluid
is slowing down when approaching
the flow front. The slight deflection in
the velocity contours is due to the transition from the Hele-Shaw to the 3-D
flow domain. While the velocities in the 3-D domain are solved using
second-order
finite difference expressions,
they are solved by Romberg
integration [64] of eqn. (10) in the Hele-Shaw domain. Grid refinement was
shown to improve the smoothness [50].
Figure 11 shows the gapwise velocity component
u in the z direction in
full view. Figure 11(b) shows the side of Fig. 11(a) in the fountain flow
region. Here, the spilling motion can be observed where fluid is flowing
towards the cavitys top and bottom walls. It is seen that the out of plane
flows are restricted
to the immediate
vicinity of the flow front, thus
justifying the previously mentioned minimum size of the 3-D flow region of
1.5 times the gapwidth. Figure 1 l(c) shows the gapwise velocity component

Fig. 11. (a) Velocity component


w for the shear-thinning
fluid (min, -0.054 m SK, max.
0.054 m SC); (b) side view of (a) at the flow front; (c) side view for Newtonian fluid (min,
-0.086 m SK; max, 0.086 m SK).

B. Friedrichs and XI. Giigeri 1 J. Non-Newtonian

Fluid Mech. 49 (1993) 141-f 73

161

0.03000
0.02500
0.02000
0.01500
0.01000
0.00500
0.00000
Fig. 12. Stream function distribution
(kin, 0.0 s-; max, 0.040 s-l).

in the Hele -Shaw domain

Fig. 13. Viscosity distribution


at the bottom
(min. 144.2 Pas; max, 14 796.9 Pas).

for the shear-thinning

fluid

half of the cavity for the shear-thinning

fluid

for the Newtonian fluid. Due to the more protruding shape of the upstream
velocity profile, the out of plane flows are more pronounced.
The maximum
value of the gapwise velocity is also seen to be located closer to the flow
front. For completeness,
the stream function distribution
in the Hele-Shaw
domain is shown in Fig. 12.

162

B. Friedrichs and S.I. Giigeri 1 J. Non-Newtonian Fluid Mech. 49 (1993) 141-l 73

The viscosity distribution


is depicted in Fig. 13 in the bottom half of the
domain. Due to extensional flow in the diverging section, the viscosity drops
along the mid-plane from the zero shear rate viscosity at the inlet. The
deformation
of fluid particles in the fountain flow region manifests itself in
a further drop close to the flow front. It is worthwhile
to note that the
steepest viscosity gradients are found in the gapwise direction. Here, the
viscosity drops by two orders of magnitude within a short distance from the
mid-plane.
This, almost shock-like
behavior,
necessitated
the use of a
conservation
formulation.
As a consequence,
satisfactory
results were
achievable using only 15 nodes in the gapwise direction.
The pressure distribution
(Fig. 14) is seen to be smooth despite the fact
that pressure and velocities were computed at the same nodal locations, i.e.
on a non-staggered
grid. Due to the continuity constraint,
the pressure is
observed not to be zero at the flow front, although the deviation is small.
In order to visualize the impact of the three-dimensional
fountain flow on
the fluid particle movement, numerical tracers were introduced at a constant
rate at the inlet. While in each batch 20 particles were introduced along the
width of the inlet, 40 were introduced through the half-gapwidth,
i.e. the top
half of the domain. These particles were relocated using eqn. (30). Since the
tracer particle locations do not need to coincide with the grid nodes, a
second-order
Taylor series expansion around the closest node was used to
interpolate for the velocities. Figure 15(a) shows the position of some tracer
particles for the shear-thinning
fluid when the simulation
was stopped.

2500000.

Fig. 14. Pressure distribution


at the bottom
(min, 134 903 Pa; max, 22 399 292 Pa).

half of the cavity for the shear-thinning

fluid

B. Fried6ehs and S.L Gii~'eri l J, Non-Newtonicm Fluid Mech. 49 (1993) 141-173

163

15
14

"K

13
12
11

[a)

~0

15

13

12
(b)

11
10

"~,

6
Fig. 15. (a) Tracer particle location at the top half of the cavity for thc shear-thinning fluid:
tracer particles were introduced slightly above the mid-plane of the cavity; (b) tracer particle
location for the Newtonian fluid.

164

B. Friedrichs and XI. Giiceri /J.

Non-Newtonian Fluid Mech. 49 (1993) 141-173

These particles were introduced at the inlet just above the mid-plane. The
numbering of the batches is done in chronological order. Therefore, batch
1 was introduced first and batch 15 last. It is seen that batch 1 has already
experienced the fountain flow while batches 2 and 3 are in the process of
doing so. The situation is much different for the Newtonian fluid. Here, more
batches have experienced the fountain flow due to the more protruding
upstream velocity profile. As a consequence, batch 1 is found further
upstream, because it has reached the fountain flow region earlier. The impact
of the diverging flow is well observed with batch 4. While the part close to
the downchannel symmetry line is only starting to feel the fountain flow, the
part at the other side has already gone through it and has, therefore, reached
the top wall. The slight waviness in batch 2 is a consequence of the velocity
interpolation. Some particles that are supposed to end up right at the top wall
do so slightly below, thus still moving with the fluid.
The situation changes when particles are traced that were introduced
halfway between the mid-plane and the top wall (Fig. 16). Here, the
shear-thinning fluid is moving faster than the Newtonian. Thus, batches 1
and 2 have experienced more of the fountain flow in Fig. 16(a) than in Fig.
16(b). This can also be observed by taking a side view along the downchannel
symmetry line of the cavity (Fig. 17). The plots are enlarged in the gapwise
direction for clarity. The characteristic V shapes formed by the individual
batches in such fountain flow situations [65] are seen to occupy different
positions and to take on different shapes due to the different constitutive
behavior of the fluids. While the V shapes formed by the Newtonian fluid are
further upstream and greater in number (Fig. 17(b)), they are stretched out
over a wider distance for the shear-thinning fluid (Fig. 17(a)). Figure 18
shows the streaklines of tracer particles introduced at 3 different positions
along the mid-plane of the cavity. Here, the three-dimensionality is perhaps
best demonstrated since particles of subsequent batches take on different
paths through the cavity in the direction perpendicular to the flow. Again,
the different fluid behaviors are seen to influence the particle relocation
dramatically.
To further demonstrate the capabilities of the solution technique the same
cavity shape was chosen at the inlet but followed by a 30 converging section.
The flow rate was lowered to 100 cm3 s- with all other input parameters
remaining unchanged. Figure 19 shows the location of the tracer particles
introduced at the mid-plane of the cavity. Again, more batches have
experienced the fountain flow in the Newtonian case than in the shear-thinning case. The fluid, however, is deformed differently as compared to the
diverging case (Fig. 15).
The computations were carried out on an IBM RS6000~550. The CPU
time was around 30 h for all the presented cases. It should be mentioned

Non-Newtonian

Fluid Mech. 49 (1993) 141 --I73

165

Fig. 16. (a) Tracer particle location at the top half of the cavity for the shear-thinning fluid;
tracer particles were introduced half way between the mid-plane and the top wall of the
cavity; (b) tracer particle location for the Newtonian fluid.

B. Friedrichs and S.I. GiiGeri1 J. Non-Newtonian Fluid Mech. 49 (1993) 141 -I 73

166

15

14

13

12

11

10

(a)

15

14

13

11

12

10

98765

(b)

Fig. 17. (a) Side view of tracer particle location along the downchannel
symmetry
the shear-thinning
fluid; (b) side view of tracer particle location for the Newtonian

line for
fluid.

that the Newtonian


case can be speeded up, because the term V * (V * t) is
zero and can therefore be dropped from the computation
of the pressure
Poisson equation (7).
8. Conclusions
A novel hybrid 2-D/3-D numerical technique was presented to model 3-D
fountain flow in thin cavities as encountered
in injection molding processes.
At the flow front, where all three velocity components
are significant, the
3-D fluid flow equations for creeping flow of a generalized Newtonian fluid
are solved using a pressure Poisson formulation.
Trailing the region, a 2-D
Hele-Shaw type formulation
is used, leading to a hammerhead
approach.
Due to the small size of the 3-D flow region, significant reduction in nodes,
i.e. unknowns,
was achieved as compared
to a fully three-dimensional
simulation. The governing equations were solved on a non-staggered
grid
using boundary
fitted coordinate
systems in conjunction
with the finite
difference
technique.
Material
non-linearities
due to the shear-thinning
behavior of typical thermoplastic
resin systems are solved for by periodically updating
the viscosity during the iteration
process. Results were

B. Friedrichs and S.I. Giigeri 1 J. Non-Newtonian

Fluid Mech. 49 (1993) 141 -I 73

Fig. 18. (a) Streaklines of tracer particles at the top half of the cavity for the shear-thinning
fluid; (b) streaklines of tracer particles for the Newtonian fluid.

167

168

B. Friedrichs and S.L Gi~geri ] J. Non-Newtonian Fluid Mech. 49 (1993) 141-173

Fig. 19. (a) Tracer particle location at the top half of the cavity for the shear-thinning fluid;
tracer particles were introduced slightly above the mid-plane of the cavity: (b) tracer particle
location for the Newtonian fluid.

B. Friedrichs and S.I. Giiceri 1 J. Non-Newtonian

Fluid Mech. 49 (1993) 141-l 73

169

presented
for the isothermal
filling of diverging and converging
cavity
sections using a shear-thinning
and a Newtonian
resin. The impact of the
3-D fountain flow on the flow field was studied by tracking numerical tracer
particles. The observed flow patterns compared well to existing two-dimensional numerical and experimental
studies.
Acknowledgments
We are grateful to Dr. S.D. Gilmore and Mr. C.A. Moore for their help
with computer graphics. This work was partially funded by the Graduate
Studies Office at the University of Delaware.
Nomenclature

w.k

B
Ca
f

g
h

iA k
J

k k,,
2

m
n
n

P, Q, R
s
St

s
t

u, 0, w
U

f.4

u
x, Y, z

covariant base vectors


multiplication
factor
concentration
factor
capillary number
field variable
diagonal components
of the contravariant
metric
gapwidth
indices
Jacobian
principal curvatures
maximum value of i
dimensionless
power law index
outward unit normal
pressure
grid control functions
incremental length
total length
fluidity
time
velocity components
in the x, y, and z directions
velocity vector
mean velocities in the x and y directions
characteristic
velocity
coordinates
in the physical domain

tensor

Greek letters
Y

surface tension
square root of l/2 times the second invariant
tensor

of the rate of strain

B. Friedrichs and S.I. Giigeri 1 J. Non-Newtonian

rate of strain tensor


viscosity
viscosity at zero shear rate
viscosity at infinite shear rate
time constant
coordinates
in the computational
density
total stress tensor
stress
stream function

Fluid Mech. 49 (1993) 141- 173

domain

References
1 L.T. Manzione (Ed.), Application
of Computer Aided Engineering in Injection Molding,
Hanser Publishers, Munich, 1987.
2 H.S. Hele-Shaw, The flow of water, Nature (London).
58 (1898) 34-36.
3 S. Richardson,
Hele-Shaw flows with a free boundary produced by the injection of fluid
into a narrow channel, J. Fluid Mech., 56(4) (1972) 609-618.
4 E. Broyer, C. Gutfinger and Z. Tadmor, A theoretical model for the cavity filling process
in injection molding, Trans. Sot. Rheology, 19(3) (1975) 423-444.
5 C.A. Hieber and S.F. Shen, A finite-element/finite-difference
simulation of the injection
molding filling process, J. Non-Newtonian
Fluid Mech., 7 ( 1980) l-32.
6 H.H. Chiang, C.A. Hieber and K.K. Wang, A unified simulation
of the filling and
post-filling
stages in injection molding. Part I: Formulation.
Polym. Eng. Sci., 31(2)
(1991) 116-124.
7 J.-L. Willien, J.-F. Agassant and M. Vincent, Numerical simulation of mold filling for
thermoplastic
injection, The Sixth Annual Meeting of the Polymer Processing Society,
PPS, Brookfield, CT. 1990.
8 P. Filz, Simulation of the injection moulding process for crosslinking materials-Current
situation and future possibilities, Kunststoffe,
79( 10) ( 1989) 1057- 1061.
9 T. Matsuoka, J.-I. Takabatake,
A. Koiwai, Y. Inoue, S. Yamamoto
and H. Takahashi,
Integrated
simulation
to predict warpage of injection molded parts., in ANTEC 90,
Society of Plastics Engmeers, SPE, Brookfield, CT, 1990, pp. 369-374.
10 P.A. Tanguy and R. Lacroix, A 3D mold filling study with significant heat effects, Int.
Polym. Proc., 6( 1) (1991) 19-2.5.
11 A. Couinot, L. Dheur, 0. Hansen and F. Dupret. A finite element method for simulating
injection molding of thermoplastics,
m J.F. Thompson
(Ed.), Numiform
89, Balkema,
Rotterdam,
1989.
12 S. Subbiah,
D.L. Trafford
and S.1. GiiGeri. Non-isothermal
flow of polymers into
two-dimensional,
thin cavity molds: A numerical grid generation approach, Int. J. Heat
Mass Transfer. 32( 3) (1989) 415-434.
13 B. Friedrichs, S. i. GiiGeri, S. Subbiah and M.C. Altan, Simulation and analysis of mold
filling processes with polymer-fiber
composites; TGMOLD.
in A.A. Tseng and S.K. Soh
(Eds.), Processing
of Polymers and Polymeric Composites,
The American Society of
Mechanical Engineers, ASME. New York, 1990, pp. 73391.
14 M.A. Garcia, C.W. Macosko. S. Subbiah and S.i. GiiGeri, Modelling of reactive filling in

B. Friedrichs and S.I. GiiGeri 1 J. Non-Newtonian

15
16
17
18
19
20
21

22
23
24
25
26
27
28
29
30
31
32

33
34
35
36
37

Fluid Mech. 49 (1993) 141-l 73

171

complex cavities: Comparison


of fountain flow approximations.
Int. Polym. Proc., 6( 1)
(1991) 73-82.
W. Rose, Fluid-fluid
interfaces in steady motion, Nature, 191 (4785) (1961) 242-243.
R.S. Spencer and G.D. Gilmore, Some flow phenomena
in the injection molding of
polystyrene,
J. Colloid Sci., 6 (1951) 1188132.
L.R. Schmidt, A special mold and tracer technique for studying shear and extensional
flows in a mold cavity during injection molding, Polym. Eng. Sci., 14( 11) (1974) 7977800.
C.G. Gogos, C. Huang and L.R. Schmidt, The process of cavity filling including the
fountain flow in injection molding, Polym. Eng. Sci., 26(20) (1986) 1457-1466.
J.M. Castro and C.W. Macosko, Studies of mold filling and curing in the reaction-injection molding process, AIChE J., 28( 2) (1987) 250-260.
H.A. Lord and G. Williams, Mold filling studies for the injection mold filling of
thermoplastic
materials, Polym. Eng. Sci., 15(8) (1975) 5699582.
S. Bhattacharji and P. Savic, Real and apparent non-Newtonian
behavior in viscous pipe
flow of suspensions
driven by a fluid piston, in Proc. 1965 Heat Transfer and Fluid
Mechanics Institute, 1965, pp. 248-262.
Z. Tadmor, Molecular orientation
in injection molding. J. Appl. Polym. Sci., 18 (1974)
1753-1772.
Manas-Zloczower,
J.W. Blake and C.W. Macosko, Space-time
distribution
in filling a
mold, Polym. Eng. Sci., 27( 16) (1987) 1229-1235.
R.C. Givler, M.J. Crochet and R.B. Pipes, Numerical prediction of fiber orientation
in
dilute suspensions,
J. Composite Mater., 17 (1983) 330-343.
D.J. Coyle, J.W. Blake and C.W. Macosko, The kinematics of fountain flow in mold
filling, AIChE J. 33(7) (1987) 1168-1177.
H. Mavridis, A.N. Hrymak and J. Vlachopoulos,
Finite element solution of fountain flow
in injection molding, Polym. Eng. Sci., 26(7) (1986) 449-454.
H. Mavridis.
A.N. Hrymak and J. Vlachopoulos,
The effect of fountain
flow on
molecular orientation
in injection molding, J. Rhoelogy, 32(6) (1988) 639-661.
R.A. Behrens, M.J. Crochet, C.D. Denson and A.B. Metzner, Transient
free-surface
flows: motion of a fluid advancing in a tube, AIChE J., 33(7) (1987) 1178-l 186.
H. Mavridis, A.N. Hrymak and J. Vlachopoulos,
Transient free-surface flows in injection
mold filling, AIChE J., 34(3) (1988) 403-410.
D. Fauchon, H.H. Dannelongue
and P.A. Tanguy, Numerical simulation of the advancing front in injection molding, Int. Polym. Proc., 6 (1991) 13-18.
R.E. Hayes, H.H. Dannelongue
and P.A. Tanguy, Numerical simulation of mold filling
in reaction injection molding, Polym. Eng. Sci., 31( 11) (1991) 842-848.
P.G. Lafleur and M.R. Kamal, A structure-oriented
computer simulation of the injection
molding of viscoelastic crystalline polymers. Part I: Model with fountain flow, packing,
solidification,
Polym. Eng. Sci., 26( 1) (1986) 92- 102.
M.R. Kamal, SK. Goyal and E. Chu, Simulation of injection mold filling of viscoelastic
polymer with fountain flow, AIChE J. 34( 1) (1988) 94-106.
AI. Isayev (Ed.), Injection and Compression
Molding Fundamentals,
Marcel Dekker,
New York, 1987.
P.J. Carreau, Rheological
equations
from molecular network theories, Ph.D. Thesis,
University of Wisconsin,
1986.
S.V. Patankar,
Numerical
Heat Transfer and Fluid Flow, Hemisphere,
Washington,
D.C., 1980.
F.H. Harlow and J.E. Welch, Numerical calculation of time-dependent
viscous incompressible flow of fluid with free surface, Phys. Fluids, 8( 12) (1965) 2182-2189.

172

B. Friedrichs and S.I. Giigeri / J. Non-Newtonian

Fluid Mech. 49 (1993) 141-l 73

38 S. Abdallah,
Numerical
solutions for the incompressible
Navier-Stokes
equations
in
primitive variables using a non-staggered
grid, II. J. Comput. Phys.. 70 (1987) 193202.
39 M.L. Mansour and A. Hamed, Implicit solution of the incompressible
Navier-Stokes
equations on a non-staggered
grid, J. Comput. Phys., 86 (1990) 147- 167.
40 K. N. Ghia. W.L. Hankey and J.K. Hodge, Use of primitive variables in the solution of
imcompressible
Navier-Stokes
equations, AIAA J., 17(3) (1979) 298-301.
41 P.M. Gresho and R.L. Sani, On pressure boundary conditions
for the incompressible
Navier-Stokes
equations, Int. J. Numer. Methods Fluids, 7 (1987) 1111-1145.
42 P.J. Roache, Short communication:
A comment on the paper Finite difference methods
by J.C. Strikwerda,
Int. J. Numer.
for the Stokes and Navier-Stokes
equations
Methods Fluids, 8 (1988) 1459-1463.
43 R.L. Panton, Incompressible
Flow, Wiley, New York, 1984.
44 B.B. Sauer and N.V. DiPaolo, Surface tension and dynamic wetting of polymers using the
Wilhelmy method: applications
to high molecular weights and elevated temperatures,
J.
Colloid Interface Sci., 144(2) (1991) 527-537.
45 E.B. Dussan, On the spreading of liquids on solid surfaces: Static and dynamic contact
lines, Annu. Rev. Fluid Mech., 11 (1979) 371-400.
46 E.B. Dussan. The moving contact line: The slip boundary condition,
J. Fluid Mech.,
77(4) (1976) 665-684.
47 C. Huh and L.E. Striven, Hydrodynamic
model of steady movement of a solid/liquid/
fluid contact line, J. Colloid Interface Sci., 35( 1) (1971) 855101.
48 H.P. Greenspan,
On the motion of a small viscous droplet that wets a surface, J. Fluid
Mech., 84( 1) (1978) 125-143.
49 E.B. Dussan and S.H. Davis, On the motion of a fluid-fluid
interface along a solid
surface, J. Fluid Mech., 65( 1) (1974) 71-95.
50 B. Friedrichs,
Modeling of three-dimensional
flow fields in injection and resin transfer
molding processes, Ph.D. Thesis, Department
of Mechanical Engineering,
University of
Delaware, 1993.
51 J.F. Thompson
(Ed.), Numerical Grid Generation,
Elsevier, New York, 1982.
52 J.F. Thompson,
Z.U.A. Warsi and C.W. Mastin, Numerical Grid Generation-Foundations and Applications,
Elsevier, New York, 1985.
53 J.P. Coulter.
Resin impregnation
during the manufacturing
of composite
materials., Ph.D. Thesis, Department
of Mechanical
Engineering,
University
of Delaware,
1988.
54 S.D. Gilmore and S.i. Gtiqeri, Solidification
in anisotropic
thermoplastic
composites.
Polym. Composites,
1 l( 6), December 1990.
55 B. Friedrichs and S.I. Giiceri. A numerical technique for the internal creeping flow of
highly shear-thinning
fluids in general curvilinear
coordinates,
Int. J. Comput.
Fluid
Dynamics,
1992, submitted for publication.
56 D.A. Anderson, J.C. Tannehill and R.H. Pletcher, Computational
Fluid Mechanics and
Heat Transfer, Series in Computational
Methods in Mechanics and Thermal Sciences,
Hemisphere.
New York, 1984.
57 J.F. Thompson,
Z.U.A. Warsi and C.W. Mastin, Boundary-fitted
coordinate systems for
numerical
solution of partial differential
equations-a
review, J. Comput.
Phys., 47
(1982) l-108.
58 S.D. Gihnore,
Thermal
and residual stress analysis in processing
of thermoplastic
composites,
Ph.D. Thesis, Department
of Mechanical Engineering,
University of Delaware. 1991.

B. Friedrichs and S.I. Giigeri / J. Non-Newtonian

Fluid Mech. 49 (1993) 141-l 73

173

59 M.N. ijzigik, Basic Heat Transfer, McGraw-Hill,


New York, 1977.
60 H. Akima, A new method of interpolation
and smooth curve fitting based on local
procedures,
J. Assoc. Comput. Machinery,
17(4) (1970) 589-602.
61 R.L. Hoffman, A study of the advancing interface, J. Colloid Interface Sci., 50(2) (1975)
228-241.
62 K.K. Wang, S.F. Shen, C. Cohen, C.A. Hieber and S. Jahanmir,
Computer-aided
injection molding system, Progress Report 5, Cornell University, College of Engineering,
Ithaca, NY, 1978.
63 K.K. Wang, S.F. Shen, C. Cohen and C.A. Hieber, Computer-aided
injection molding
system, Progress Report 6, Cornell University, College of Engineering, Ithaca, NY, 1979.
64 W.H. Press, B.P. Flannery, S.A. Teukolsky and W.T. Vetterling, Numerical Recipies,
Cambridge University Press 1986.
65 A.N. Beris, Fluid elements deformation
behind an advancing flow front, J. Rheol., 31(2)
(1987) 121-124.

Vous aimerez peut-être aussi