You are on page 1of 13

.

Powder Technology 106 1999 119131


www.elsevier.comrlocaterpowtec
Unsteady state planar divergent flow of extrusion pastes
A.S. Burbidge
a,)
, J. Bridgwater
b
a
School of Chemical Engineering, Uniersity of Birmingham, Edgbaston, Birmingham B15 2TT, UK
b
Department of Chemical Engineering, Uniersity of Cambridge, Pembroke Street, Cambridge CB2 3RA, UK
Received 6 November 1998; accepted 29 March 1999
Abstract
The divergent downward flow of pastes between parallel plates is examined experimentally and several distinct phases are identified in
the development of the flow. The experimental design allows the distance between the plates to be adjusted. Initially, a bar of paste is
formed but forces arising from the sliding contact cause a back pressure on the feed nozzle which leads to bulging of the paste close to
the entry port. This observation allows an alternative method of determining paste properties which potentially gives better estimates of
the bulk yield stress than those found by a piston extrusion method. In subsequent phases, the flow is radial when close to the feed but
becomes more complex at greater distances. A radial flow model is derived and compared with experimentally determined values of the
radial stress component at the inlet. In some cases, the agreement is good. However, the paste does not always flow radially and the
surface contact with the plates is not always good. The flow fields should prove valuable in the development of numerical models; these
also show how defects can develop in products made by moulding. q1999 Elsevier Science S.A. All rights reserved.
Keywords: Unsteady state; Divergent flow; Extrusion pastes
1. Introduction
Extrusion is a widely used process of considerable
commercial importance. The products fabricated in this
way range from polymers to ceramics, foods and metals.
Despite the importance of extrusion as a process, little is
known about the flow of extrusion pastes. The advent of
injection moulding for paste systems means that this lack
of knowledge needs to be addressed.
A typical extrusion paste is a dense, polydisperse sus-
pension of solid phase particles in a lubricating liquid
phase. The solid phase may be, say, a ceramic particle of
.
small diameter typically -50 mm , although it is com-
mon to improve the packing fraction by adding smaller
particles that can partially fill the voids between those of
larger size. The liquid phase is commonly an aqueous
suspension of a rheology modifier, the typical ones being
clays, carbohydrates and water soluble polymers such as
methyl-celluloses. It is important to choose a liquid phase
)
Corresponding author. Tel.: q44-121-414-5295; fax: q44-121-414-
5377; E-mail: a.s.burbidge@bham.ac.uk
carefully so as to produce a consistent homogeneous prod-
uct and to avoid quality control problems.
In a paste flow, there is a complex interaction between
the liquid and solid phases. Liquid phase movement within
a paste prevents the use of conventional oscillatory cone
and plate rheometry, since any sample of paste will soon
generate lubricating layers at the shearing surfaces. Oven-
w x
ston and Benbow 1 suggested that any changes in cross-
section could be approximated by using the pure plastic
work assumption common in metal deformation theory.
They combined this with the assumption that in a pipe of
constant cross-section, there is a plug flow with a thin
layer of lubrication at the wall. Hence, they present a
semi-empirical model relating pressure drop to the geome-
w x
try of a ram extruder system. Benbow and Bridgwater 2
present this model in detail and describe how it can be
adapted for systems other than flow from a barrel into a
die in a ram extruder. Although the flow of extrusion
pastes can be described semi-empirically for axi-symmet-
ric flows, the extension of models to non-symmetric sys-
tems is problematical at the present time. The present
purpose is to develop an understanding of paste injection
between parallel plates. Vertical downflow is studied, the
0032-5910r99r$ - see front matter q 1999 Elsevier Science S.A. All rights reserved.
. PII: S0032- 5910 99 00069- 8
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 120
Fig. 1. Laboratory piston extruder.
purpose being to gain an understanding of the flow, to
explain it quantitatively and to discover the origin of
defects in structure.
2. Paste preparation
Three pastes were used in this study: paste A was
.
composed of 25.5% wt. each of 3, 17, 37 mm median
.
diameter a-alumina particles, 3% wt. each of potato
.
starch and bentonite, and 17.5% wt. water. Paste B was a
.
boehmite gel composed of 50.5% wt. boehmite, 46.5%
. .
wt. water and 3% wt. HNO . Paste C was composed of
3
.
27.3% wt. each of 3, 17, 37 mm median diameter a-
.
alumina particles, 4% wt. of potato starch and 14.1%
.
wt. water. Alumina was obtained from Universal Abra-
sives, Bentonite from Steetly Minerals. Potato starch was
.
supplied by Sigma and boehmite Plural SB from Condea
Chemie. In preparation of A and C, the dry components
were mixed together for 20 min in a Hobart mixer and
Fig. 2. Parallel plates apparatus.
Fig. 3. Elemental force balance for a small radial flow element.
then the water was added before a further 20 min mixing.
The paste was then passed twice through the mincer
attachment of the mixer to remove agglomerates. To make
paste B, the acid was diluted with the water, this added to
the boehmite powder and the whole lot mixed in a Z-blade
mixer for 20 min. All pastes were stored in a refrigerator
in PVC bags to prevent loss of liquid phase by evapora-
tion. The paste was allowed to return to room temperature
before experiments were performed.
3. Experimental method
Experiments were conducted in a modified piston ex-
.
truder system Fig. 1 constructed around a Dartec com-
puter-controlled 100 kN screw strain frame. The cross-head
of the strain frame was connected to the plunger of a
piston extruder constructed of stainless steel. A ribbed
nylon seal ensured that there was no leakage of the sample
at the top surface. The plunger was closely fitting inside a
25-mm barrel section that screwed into an angled flange
which connected self-centering die attachments. The entire
piston arrangement was supported by means of a four-
legged stand constructed of mild steel, allowing access to
the outlet of the die attachment. For the experiments
detailed here, a parallel plate apparatus was attached to the
.
piston extruder Fig. 2 . This consists of a pair of 100
mm=100 mm=10 mm polycarbonate plates mounted
such that the separation can be adjusted between 1 and
Fig. 4. Mohrs circle of stress for small elements of paste at different
radii.
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 121
20 mm. The width of the slit in the plane of the plates was
fixed at 20 mm. A Panasonic SVHS video camera could be
positioned normal to the plane of the plates was used to
provide a visual record of the extrusion flows.
The barrel of the extruder was filled with paste and the
plunger positioned so that the cross-head of the strain
frame was in direct contact with the upper surface of the
paste in the barrel. This was taken as the reference for
displacement. The extrusion program was then executed,
operation being completely automatic until the required
extrusion had been completed. The cross-head was then
backed away from the top of the paste, so that it was
possible to remove the extruder assembly from the strain
frame.
The experiments were videoed and images processed
using a PC with a frame grabber card. Images could thus
be automatically acquired at a range of time intervals over
the duration of the experiment. The resulting sequences of
images could then be viewed either individually or as a
multimedia movie. Strain frame data was also available
describing displacement and extrusion pressure vs. time.
w x
Benbow and Bridgwater 2 rheological parameters for
the extrusion pastes were measured in the standard way.
This involved extrusions through square entry dies of
Fig. 5. Comparison of extrusions of pastes A, B and C between parallel plates of pitch 10 mm at creeping flow rates.
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 122
Fig. 6. Force balance on initial extruded finger of paste at point of yield.
varying lengths at a range of velocities. The parameters are
recovered by means of a set of least squares regressions.
4. Modelling
Consider an element of paste of angle f and thickness
.
w Fig. 3 . Define a coordinate frame by the unit vectors
e in the radial direction, e in the tangential direction and
1 2
e normal to the plane of the plates. Further, take the sign
3
convention that compressive stresses are positive and as-
sume a flowfield such that there is uniform velocity in the
radial direction of magnitude independent of the angular
coordinate.
Table 1
BenbowBridgwater extrusion parameters for pastes A, B and C mea-
sured by method of finger yielding between parallel plates
. . . Paste Plate pitch s MPa t Mpa b MPa srm
0 0
. mm
A 4 0.18 0.01 0.00
A 10 0.30 0.02 0.03
A 15 0.22 0.02 0.11
B 4 0.01 0.00 0.00
B 10 0.29 0.01 1.26
B 15 0.31 0.08 0.00
C 4 1.84 0.04 0.00
C 10 1.56 0.08 0.00
Consider the geometry of an arc at radius r, the differ-
ential annular surface area A normal to the 3 direction,
3
i.e., in the plane of the paper is
f
2
2
d A s rqdr yr (frdr 1 . .
3
2
and the flow area normal to radial direction at radius r is
A sfrw 2 .
1
Hence, assuming no velocity in the 3 direction, the superfi-
cial radial velocity is given by
Q
s 3 .
1
wrf
since the material is incompressible. For a paste system
Fig. 7. Typical forcerdisplacement data from Dartec strain frame illustrating calculation of net piston force acting on extruded finger of paste paste A,
. 4-mm gap, creeping flow .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 123
Table 2
BenbowBridgwater extrusion parameters for pastes A, B and C mea-
sured by piston extrusion method
. . . . Paste s MPa a MPa srm t MPa b MPa srm
0 0
A 0.33 1.13 0.02 0.27
B 0.43 0.11 0.05 0.70
C 1.86 38.3 0.04 2.94
w x
sliding on a surface, Benbow and Bridgwater 2 present
.
the following result z)0 :
t
0 i
t s qb 4 .
i i
< <
z
where t is the ith component of a lubricated surface shear
i
stress vector t due to the slip velocity vector z of the
paste at a point. t is the stress required to initiate slipping
0
and b is a velocity factor. This equation amounts to
allowing a thin lubricating layer of Bingham fluid at the
surface of a bulk paste flow which is undeformed normal
to the plane of slip. Since the thickness of this shear layer
is unknown, it is more convenient to work with slip
velocities than shear rates.
The rate of work input at the inner surface normal to the
radial direction is
w ss
rsr
0
wr ss
rsr
0
Q 5 .
r 11 1 11
where s
rsr
0
is the radial normal stress component mea-
11
sured on the inner face of the radial flow region. The rate
of work done by surface shear stress on one pasterplate
surface is given by
a
w s frt dr 6 .
H
s 1 1
r
0
the components in directions 2 and 3 being zero since
s s0. This can be evaluated by substituting from 3
2 3
into 4 to give
a a
2
Q bQ
w s t drq dr
H H
s 0 2
w frw
r r
0 0
Qt bQ
2
a
0
w s ayr q ln 7 . .
s 0 2
/
w r fw
0
The internal dissipation rate D
U
within the volume V
associated with an arbitrary strain rate field
U
is

U U
D s s : dV 8 . H
V
where s is the stress field satisfying the yield criterion
over the volume V, which, for a two-dimensional system
along arbitrary principle directions UU and VV, reduces to

U
D s s qs dV 9 . . H
UU UU VV VV
V
.
For an incompressible material tr s0, i.e., q
UU
.
s0. Substituting into Eq. 9 yields
VV

U
D s s ys dV 10 . . H
UU VV UU
V
For a perfectly plastic material obeying Trescas yield
Fig. 8. Typical divergent flow pattern showing two-dimensional nature of the flow pattern in the plane of the plate surfaces paste A, 10 mmrs, 10-mm
. plate spacing .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 124
. Fig. 9. Creeping flow summary for paste A 1 mmrs superficial velocity at entry .
criterion s ys ss for which is by definition
UU VV 0 UU
the maximum principal strain rate further reduces to

U U
< <
D s s dV 11 . H max
0
V
which is a well known plasticity result. The components of
the strain rate tensor in cylindrical polar coordinates are
w x
given in many standard texts, e.g., Bird et al. 4 . For our
assumed strain field where s0 this results in
u
E Q
r
s sy 12 .
11 2
E r fr w
1 E Q
u r
s q s 13 .
22 2
r Eu r fr w
E 1 E
u r
sr q s0 14 .
12
/
Er r r Eu
where s since the strain rate tensor is symmetric.
12 21
Hence, directions 1 and 2 are principal axes and thus
11
and are equal and opposite maximum strain rates. It
22
should be noted that for this to be true, the shear rate in the
13 directions needs to be vanishingly small. In the case
.
of the experiment, this is approximately satisfied Fig. 8 .
. .
Thus, from Eqs. 11 and 12
E Q Q
U
< <
s s 15 . max
2
/
E r frw fr w
so
Q
U

D s s dV 16 .
H
0 2
fr w
V

2
.
Noting Vs fr wr2 . Changing the variable from V to r
and integrating between outer radius a and inlet radius r
0
yields
a s Q a
0
U

D s drss Qln 17 .
H
0
/
r r
r
0 0
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 125
An energy balance on the system equating the total work
.. ..
input Eq. 5 with that output Eq. 6 or dissipated
..
internally Eq. 17 leads to
a 2t 2bQ a
0
rsr
0
s ss ln q ayr q ln .
11 0 0 2
/ /
r w r fw
0 0
18 .
In the case of an expanding disc of material, the value of a
is then a function of time, and can therefore be substituted
with a function of Q and elapsed time, t, by means of a
mass balance.
.
When the substitution fs2p is made, Eq. 18 is
w x
identical to a result presented by Bates 3 for radial
outflow between a pair of discs. His approach was based
upon integration of a differential force balance.
Consider the stress state at various radii within a disc of
. .
paste Fig. 4 . Case a represents the stress state at the
outer edge of the disc of paste. A necessary boundary
condition is that the stress normal to the surface is zero
and this fixes the point s at the origin of Mohrs circle
11
of stress. Assuming that the material is in plastic yield,
Trescas criterion fixes the circumferential stress s at a
22
value ys . Now consider a small element at some smaller
0
radius such that radial stress s is now compressive. It
11
can be seen that the hoop stress s becomes less and less
22
.
tensile until finally at a neutral point b it is zero. At radii
less than this neutral point, both stress components are
.
compressive c .
If the yield criterion is symmetric about the shear stress
axis, i.e., the behaviour of the paste is identical in tension
and compression, then the change from tensile to compres-
sive hoop stresses is mechanistically unimportant. How-
ever, if the paste has significantly less strength in tension
than compression, its behaviour is more complex a
matter addressed below.
5. Results and discussion
Fig. 5 shows images captured at various stages during
typical extrusions of pastes A, B and C. In each frame, the
bar that appears along the top of the image is a 100-mm
long brass beam which is an integral part of the frame and
. Fig. 10. Fast flow summary for paste A 10 mmrs superficial velocity at entry .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 126
is external to the plates, and thus has no effect on the
observed flow field. The 10 mm=100 mm square dieland
channel extends to the level of the lowest part of this
support and hence the view of the divergent region is
uninterrupted. The bolts that appear in the lower right and
left corners of the frame hold the plates parallel and at the
required pitch. Each bolt ensures a tight contact between
the plates and a spacer which is sandwiched in-between.
Rectangular spacers maintain the pitch of the plates at the
injection point at the top of each image and also form the
edges of the dieland. The images shown here are for a
plate spacing of 10 mm. The width of the injection point
was thus held constant at 10 mm throughout, so in the case
of these images, the injection point was a square cross-sec-
tion of 10 mm=10 mm. In these experiments, the superfi-
cial paste velocity through this orifice was maintained at 1
mmrs.
In each case, during the initial stages of the flow, the
paste forms a finger of the same cross-sectional area as the
.
injector Fig. 5a,b,c . This finger will eventually form the
stalk of a mushroom shaped extrudate. At this stage,
there is no divergence of the flow pattern since the stress
state of the paste at the entry to the parallel plates is less
than the bulk yield stress. As the outflow continues, the
contact area of the paste and plate becomes larger; there-
fore, the stress at the entry due to the surface shear stress
increases. As the yield stress is approached, the finger
begins to bulge until ultimately there is a change in the
observed flow pattern. In the case of pastes A and B, a
.
mushrooming occurs Fig. 5d,e , whereas for paste C the
flow splits to one side and the other of the initial finger
which is now stationary. Following a mushroom analogy
as before, this bulging region forms the cup of the
mushroom. The flow patterns of pastes A and B approach
the expected radial flow pattern predicted by the theory,
although cracks appear at the top of the stalk. As the
paste is undergoing a hoop stretching, there will be a
tensile hoop stress acting along the free surface of the
mushroom. It seems likely that the junction of this expand-
ing outer surface and the finger originally formed act as
. Fig. 11. Creeping flow summary for paste B 1 mmrs superficial velocity at entry .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 127
stress concentrators causing preferential cracking. Paste C
.
is behaving differently Fig. 5f and is not consistent with
a radial flow pattern. Rather a sequence of fingers is seen.
After some time, the flow patterns of pastes A and B
settle down into a region of radial flow with a collection of
cracked peninsulae at the periphery. It is notable that the
cracks that were apparent at small radii at short times have
either healed or convected with the flow, since they are no
longer visible as the experiment progresses. This behaviour
is consistent with the radial flow model since the hoop
stress will remain at a maximum at the free outer surface
and become more compressive at lower radii. Assuming
these are indeed caused by hoop tension, their positions
will therefore be determined by the distance from the outer
surface. This would seem to be consistent with convection
.
of cracks. In the case of paste C Fig. 5i , there would
seem also to be radial flow surrounded by a cracked outer
region, although in this case, the radial flow region is very
small. It seems in this case that the tensile yield stress of
the material is almost negligible when compared with that
in compression. If this were the case, then radial flow
could only occur at radii small enough that the hoop stress
.
was compressive Fig. 4 . This would account for the
small radial flow region.
Since the initial stages of the flow lead to the extrusion
of an undeformed finger of material, from statics, a force
.
balance can be attempted Fig. 6 . Assuming that there is
good contact between the plates and the surface of the
paste allows calculation of values for the surface shear
stress t , the surface shear stress velocity factor b and the
0
bulk yield stress s . The force on the extrusion piston is
0
.
measured as a function time Fig. 7 ; hence, the fraction of
the force acting on the finger of interest can be calculated
by subtraction of a reference pressure noted at the point
the finger begins. In these experiments, the contribution of
the self-weight of the paste is negligible being less than
1% of the bulk yield stress in all cases, and is therefore
ignored in the calculations. If, for the creeping flow exper-
iments, it is assumed that the flow rate is small enough that
the contribution of the b term can be neglected, then
. .
s s Frwt and t s Fr2wL where L is the length
0 0 1 1
at which the finger first starts to bulge. Consideration of a
fast flow experiment also at superficial velocity V leading
2
to an initial bulging length of L allows calculation of b
2
. Fig. 12. Fast flow summary for paste B 10 mmrs superficial velocity at entry .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 128
. .
values since b' L t yL t r L V . Values of s , t
1 0 2 0 2 2 0 0
and b calculated by this method are shown in Table 1. In
subsequent text, values calculated from these parallel plate
experiments will be referred to as s
U
, t
U
and b
U
so as to
0 0
distinguish them from values calculated by the standard
w x
piston extrusion method Ovenston and Benbow 1 , Table
.
2 .
For paste A, the values of s
U
are consistently lower
0
than s values measured by the piston technique. Values
0
of t
U
are in agreement with t for the 10 and 15 mm
0 0
spacings although the value calculated from the 4 mm
spacing is too low. All values of b
U
are far too small, the
closest approach being for a 15-mm spacing giving a value
of 0.11 MPa srm compared with the 0.27 MPa srm from
the capillary experiment.
For paste B, the value of s
U
is very low for 4-mm
0
plates, although for 10- and 15-mm plates, the estimate is
about 25% lower than that from the standard method.
Values of t
U
are in fair agreement for 10 and 15 mm
0
spacings although the value calculated from the 4-mm
spacing is too low. Comments on values of b
U
are not
significant given the degree of scatter. It should be noted
that for paste B the length of the unyielded finger is very
short leading to a larger error in these measurements than
that found for pastes A and C.
In the case of paste C, the values of s
U
are in excellent
0
agreement with those values measured by the standard
method for both spacings tested. t
U
estimates are also in
0
fair agreement, although values of b
U
are once more
unreliable. Results are not included for the 15-mm plate
since the apparatus was not sufficiently long to cause the
finger to yield before it left the system at the bottom edge
of the plates.
This parallel plate method would seem to provide a
lower value of the bulk yield stress than the extrusion
method. This is not surprising, since the extrusion method
assumes that the plastic deformation due to the change of
flow area is completely uniform and that the work done is
equivalent to the perfect plastic work of the deformation.
In fact, the deformation is almost certainly non-uniform
leading to extra or redundant work being required. This
redundant work will cause the value of bulk yield stress
calculated by this method to be an overestimate. If there is
no extrudate swelling, then the parallel plate method pro-
. Fig. 13. Creeping flow summary for paste C 1 mmrs superficial velocity at entry .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 129
vides a pure compressive stress at the top of the extruded
finger and would therefore be expected to give a better
estimate of the bulk yield stress. Measurements of the
shear stresses t
U
and b
U
are dependent upon the reliabil-
0
ity of the surface contact and therefore prone to error when
there is a low back-pressure. It would thus be expected that
these parameters measured from the capillary experiment
would provide a better estimate.
A feature of the flow is its extreme two-dimensionality
.
in the plane of the plate surfaces Fig. 8 . Note particularly
the extremely sharp square edge at each of the plate paste
interfaces and the straightness of the free surface joining
the plates. In the case of a typical Newtonian fluid, this
would be a parabolic surface, but the surface observed is
consistent with a lubrication assumption as was used in the
derivation of the radial flow model. Defects in the paste
run from one plate to the other. There is negligible die
swell in these experiments.
A crude test of the radial flow theory can be made by
measuring the pressure at the injection point as a func-
tion of time. From a knowledge of the flow rate the
apparent radius a can be calculated if it is assumed that the
material flows radially at all times. We know that notably
for pastes A and B the paste does not flow radially at early
times, although as the experiment progresses the flow
becomes increasingly radial in nature and a then becomes
a better characteristic length scale. The extrusion force on
the piston can be measured as previously detailed, allow-
ing the calculation of a pressure at the injection point of
the plates. There is a degree of uncertainty in associating
this pressure with the stress components predicted by the
model, since the edge of the injection slot does not align
conveniently with the cylindrical polar coordinate frame
employed in the model. That is, we are trying to relate the
measurement of an axial stress to a predicted radial stress.
This issue can be addressed by carrying out a force
balance on a control volume chosen to totally enclose the
piston and have a bottom boundary that coincident with a
5-mm radius semicircle that just touches either side of the
injection port. Hence, this radial normal stress component
is lower than the axial normal stress calculated along the
physical edge of the inlet by a factor of pr2. Experimental
radial stresses were all calculated in this manner, and these
represent the normal stresses at the 5-mm semicircle.
. Fig. 14. Fast flow summary for paste C 10 mmrs superficial velocity at entry .
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 130
Figs. 914 and Fig. 15 show comparisons between
experimental radial stress components as determined by
this method and predictions from the radial flow theory.
The rheological parameters used in the theory were those

determined by the standard piston extrusion method Table


.
2 . In all cases, the stress data has been normalised so that
when as0.005 m, the radial normal stresses are zero, but
no other adjustments are made. This can be tricky in
interpreting some experiments, the results of this being
particularly prevalent in the case of the larger plate gaps.
For example, see Fig. 9 with a 15-mm gap where at later
times the forms of the theoretical and experimental curves
are very similar, although the absolute values are displaced
due to anomalous behaviour at the start of the experiment.
For paste A, experimental results for creeping flow
.
seem to show good agreement with theory Fig. 9 , al-
though the theory is an overestimate for larger gaps. This
can probably be attributed to the normalisation procedure
discussed in the previous paragraph and also to a reduction
in the contact at the interface of the plates and paste. For
.
fast flow Fig. 10 , the 4-mm gap data again show the
correct functional behaviour although the absolute value of
the theory is too low. The theoretical gradients for the
larger gaps are overestimated. It is unclear why this effect
is seen at higher superficial flow rates.
.
Experiments with paste B Figs. 11 and 12 show very
poor agreement with theory beyond radii of 0.02 m. In all
cases, the experimentally measured stresses become con-
stant or start to fall, while the theoretical predictions
continue to rise. Again, these effects can almost certainly
be attributed to contact problems at the sliding paste
surface. Paste B is a gel and hence has considerable
elasticity and therefore one could anticipate a Poisson
contraction normal to the plane of straining, this contrac-
tion reducing the contact of paste and plate and therefore
lowering the observed stress gradient. Such effects are not
included in the current model since perfect plasticity is
assumed with elastic strains being zero. This mechanism of
Poisson contraction had been suggested previously by
w x
Bates 3 .
For completeness, we consider finally the experimentr
.
theory comparisons for paste C Figs. 13 and 14 . Gener-
ally, there is little sign of agreement except in the case of
.
fast flow with a 10-mm gap Fig. 14 . This is probably
nothing more than coincidence since in all four cases paste
clearly does not flow radially and hence the radial flow
theory is not applicable.
6. Conclusions
Experiments have been performed on pastes between
parallel flat plates of varying gaps with three different
pastes. The development of the flow patterns and injection
pressures with time have been recorded. In each case, the
flow patterns are found to be controlled mainly by the
formulation of the pastes: one exhibiting radial flow; a
second mushrooming; and a third fingering. This fingering
is thought be due to a significantly reduced yield stress in
tension. Paste properties such as tensile yield stress appear
to be extremely sensitive to formulation. In order for a
paste to flow in a divergent manner, it must be able to
support a hoop tension; therefore, formulation of pastes for
moulding is a very important consideration.
A consideration of the early stages of flow allows an
estimation of the compressive bulk yield stress, the surface
shear yield stress and the surface shear stress velocity
factor of the flowing material. Comparisons with measure-
ments made by the conventional piston extruder method
w x
due to Ovenston and Benbow 1 and Benbow and Bridg-
w x
water 2 lead to the conclusion that this plate method
provides a potential for an improved measurement of s ,
0
but is less reliable for measuring the surface characteristics
t and b.
0
A radial flow model has been derived for a segment of
lubricated, perfectly plastic paste in radial outflow. The
result has been shown to be identical to that previously
w x
presented by Bates 3 , although it is derived by a consid-
eration of work and not a force balance. Comparisons
between the predicted and experimentally determined val-
ues of radial stress at the inlet have been made for three
different pastes, three different plate spacings at both
creeping and fast flow rates. The radial flow theory seems
to generally produce an overestimate of the measured
stress values, the degree of overestimate depending on how
close the experiment approaches radial flow. Fast flow
values are also less good estimates than those for creeping
flow, suggesting that the surface shear values are not
correct. This may be due to a Poissons ratio effect reduc-
ing the contact between the plates and the paste.
The work gives rise to the need for further studies in
several areas. The next step will be to vary the angle of
segment by further constraining the flow and compare
theory and experiment, this being the steady-state experi-
ment in contrast to those of unsteady state detailed above.
Ultimately, this idea will be extended to a consideration of
systems with parallel restraining walls and very narrow
separations as found in honeycomb die systems. Further
into the future, flow tracking techniques will be developed
in order to better understand the flows in this twin plate
system. Additionally, the effects of injection velocity and
plate spacing will be further investigated. Attempts are
also underway to measure independently the tensile yield
properties of extrusion pastes, a necessary corollary of
established methods employed in compression.
7. Nomenclature
.
a Outer radius of paste segment m
A Surface area of paste segment in plane normal
x

2
.
to x direction m
( ) A.S. Burbidge, J. BridgwaterrPowder Technology 106 1999 119131 131

U
.
D Rate of internal work dissipation Jrs
e Unit vector in i direction
i
.
F Net force acting on paste finger MN
L Length at which extruded finger begins to
1
.
bulge during creeping flow m
L Length at which extruded finger begins to
2
.
bulge during fast flow m

3
.
Q Volumetric flow rate of paste m rs
.
r Radius m
.
r Inner radius of paste segment injector radius
0
.
m
.
t Width of injector m

3
.
V Volume of paste segment m
.
V Velocity of paste in x direction mrs
x
.
w Thickness of paste sample m
.
w Work done by injection stress J
r
.
w Surface shear work J
s
.
b Surface shear stress velocity factor MParm
b
U
Surface shear stress velocity factor estimated
.
from plates experiment MParm
U

y1
.
Assumed strain rate field tensor s
.
f Angle of paste segment rad
.
s Stress tensor MPa
.
s Bulk yield stress MPa
0
s
U
bulk yield stress estimated from plates experi-
0
.
ment MPa
.
s i, j component of stress tensor MPa
i j
s
rsr
0
Radial normal stress component at injector
11
.
MPa
t Component of surface shear stress vector in i
i
.
direction MPa
.
t Shear yield stress of surface layer MPa
0
t
U
Shear yield stress of surface layer estimated
0
.
from plates experiment MPa
Acknowledgements
Thanks are due to both A. Bates and D. Horrobin for
discussions regarding the modelling and to Z. Saracevic
for assistance with experimental work. F. Kolenda at the
Institut Francais du Petrole, Solaize kindly supplied us
with some of the experimental materials. This work was
carried out under an EPSRC Soft Solids grant.
References
w x . 1 A. Ovenston, J.J. Benbow, Trans. Br. Ceram. Soc. 67 1968 543.
w x 2 J.J. Benbow, J. Bridgwater, Paste Flow and Extrusion, Oxford Univ.
Press, Oxford, UK, 1993.
w x 3 A.J.D. Bates, PhD Thesis, Univ. of Birmingham, Birmingham, UK,
1995.
w x 4 R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena,
Wiley, New York, 1960.

You might also like