Sunteți pe pagina 1din 10

Effects of polymer melt compressibility on mold filling in micro-injection molding

This article has been downloaded from IOPscience. Please scroll down to see the full text article.
2011 J. Micromech. Microeng. 21 095019
(http://iopscience.iop.org/0960-1317/21/9/095019)
Download details:
IP Address: 155.69.203.3
The article was downloaded on 19/08/2011 at 02:39
Please note that terms and conditions apply.
View the table of contents for this issue, or go to the journal homepage for more
Home Search Collections Journals About Contact us My IOPscience
IOP PUBLISHING JOURNAL OF MICROMECHANICS AND MICROENGINEERING
J. Micromech. Microeng. 21 (2011) 095019 (9pp) doi:10.1088/0960-1317/21/9/095019
Effects of polymer melt compressibility on
mold lling in micro-injection molding
Q M P Nguyen
1
, X Chen
1
, Y C Lam
1,2
and C Y Yue
1,2
1
Singapore-MIT Alliance, Nanyang Technological University, 65 Nanyang Drive, Singapore 637460
2
School of Mechanical & Aerospace Engineering, Nanyang Technological University,
50 Nanyang Avenue, Singapore 639798
E-mail: myclam@ntu.edu.sg
Received 14 April 2011, in nal form 24 June 2011
Published 18 August 2011
Online at stacks.iop.org/JMM/21/095019
Abstract
In conventional injection molding, the molten polymer in the lling stage is generally assumed
to be incompressible. However, this assumption may not be valid in micro-injection molding,
since high injection pressure is normally required to avoid short shots. This paper presents
both numerical and experimental investigations on the effects of polymer melt compressibility
on mold lling into a micro-thickness impression. The study was conducted on six different
part thicknesses ranging from 920 to 370 m. A high-ow COC TOPAS 5013L-10 polymer
was chosen as the TOPAS family has recently attracted signicant interest for its use in
microuidic applications. A combined nite element/ nite difference/ control volume
approach was adopted to simulate the compressible ow. The shear viscosity of a polymer
melt was characterized by the Cross-WLF model, while the melt compressibility was modeled
with a double-domain Tait equation. The results obtained indicated that the compressibility of
the polymer melt has signicant effects on impression pressure and density distribution in the
fully lled part with thickness smaller than 620 m and that the effects become more
pronounced with a decrease in part thickness.
(Some gures in this article are in colour only in the electronic version)
1. Introduction
Due to its high productivity and high exibility, micro-
injection molding is the preferred process for producing large
volume micro-plastic parts at low cost and short cycle times.
A typical injection molding cycle involves a series of distinct
stages, namely lling, post-lling and cooling stages. A good
ow pattern within the impression during the lling phase is
an important consideration for a successful molded part. As
such, it is advantageous to predict accurately the owbehavior
to reduce or to avoid the expensive iterations in physical mold
trials.
Extensive numerical simulations on mold lling have
been carried out by various research groups. A considerable
advance in ow analysis was made by Hieber et al [1],
who rst employed the hybrid nite element/nite difference
scheme for a generalized Hele-Shaw ow in a thin impression
under non-isothermal conditions. By combining this hybrid
formulation with the ow analysis network (FAN) [2],
Wang et al [3] introduced the control volume approach to
capture the ow front in three-dimensional geometries. It is
noteworthy that the combined nite element/ nite difference/
control volume approach has become the standard numerical
framework for numerous in-house research codes as well as
commercial simulation packages [47].
With increasing demands for high-quality products and
to minimize manufacturing costs, commercial ow analysis
software has become a routine tool for designing plastic
parts. However, simulation packages that can successfully
predict the ow pattern in macro-scale injection molding
might not provide accurate predictions for micro-injection
molding for the whole range of processing conditions [810].
The discrepancies between numerical and experimental results
could be caused by (a) micro-scaled factors such as wall slip,
surface tension, surface roughness, etc, which are generally
neglected and not important in conventional injection molding,
(b) the lack of good quality material databases, and (c) the
neglect of polymer melt compressibility in the lling phase.
0960-1317/11/095019+09$33.00 1 2011 IOP Publishing Ltd Printed in the UK & the USA
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
Melt compressibility has effects on both the polymer
density and the pressure distribution within an impression. The
effects of melt compressibility are generally negligible for the
melt ow at low or moderate pressure normally experienced
in macro-injection molding. However, with a decrease in
dimensions of micro parts, relatively high injection pressure
and high injection rate are required to avoid short shots
[11, 12]. Thus, the compressibility of the molten polymer may
have signicant effects on the ow behavior, on shrinkage and
warpage as well as on the replication of micro-features [13].
It should be noted that in most numerical simulations
for both macro- and micro-injection molding, the material
is assumed to be incompressible during the lling phase
but compressible in the post-lling stage. As such, the
melt compressibility effects on mold lling are neglected
and a separate set of governing equations are required for
each stage. A few studies were conducted previously on
the compressible ow in the lling stage. Back in the
early 1990s, Chiang et al [14, 15] developed a unied
numerical analysis for both lling and post-lling stages
by employing the same theoretical model for both stages.
Experimental verications of their model were conducted
for both amorphous and semi-crystalline polymeric materials.
Although the continuity equation for a compressible uid was
used in their investigation, it was found that the assumption of
a constant density (i.e. the incompressible ow) is adequate for
the impression-lling analysis in the macro-injection molding
process. Following this approach but incorporating the
phase change effect, Changyu et al [16] reported that the
compressibility of the molten polymer should be considered
in the lling stage of an unbalanced mold. For micro-injection
molding, the compressible lling ow into micro-spiral and
micro-plate impression was considered by Tham et al [17]
with pressure-dependent melt viscosity and by adjusting heat
transfer coefcients in the commercial Moldow Plastics
Insight. However, there were no clear conclusions on the
melt compressibility effects in their study.
By developing our in-house ow analysis in C language
which adopts a 2.5D coupled Hele-Shaw approximation
for the compressible viscous ow under non-isothermal
conditions, this investigation presents a systematic numerical
and experimental investigations on the compressibility effects
of the molten polymer in the lling of a micro-thickness
impression.
2. Theory
For convenience, a Cartesian coordinate system is used to
describe the impression as shown in gure 1. Here the molten
polymer is occupying a three-dimensional region (t) at time t.
C
e
is the entry contour through which the melt enters the
impression and C
m
(t) is the surface dening the moving front.
The outer impression boundary C
o
includes the top and the
bottom surfaces of the mold wall and the surface dening
the edge of the mold which is in contact with the melt. The
gap thickness of the part 2h is assumed to be much smaller
than the length scale dening C
o
. This assumption is valid
for the simulation of micro-injection molding since a micro
Figure 1. Schematic diagram for mold impression.
part generally has a thickness/width ratio of much less than
1. Finally, denes the location of the solidliquid interface.
2.1. Governing equations
The motion of a polymer melt is governed by the principles
of conservation of mass, momentum and energy. For a
generalized compressible inelastic Hele-Shawowunder non-
isothermal conditions, this set of equations can be written as
[5, 14, 18]

t
+
(u)
x
+
(v)
y
+
(w)
z
= 0 (1)
p
x
=

z
_

u
z
_
(2)
p
y
=

z
_

v
z
_
(3)
C
pl
_
T
t
+ u
T
x
+ v
T
y
_
= k
l

2
T
z
2
+
2
(4)

s
C
ps
T
s
t
= k
s

2
T
s
z
2
. (5)
Here is the density, u, v and w are the velocity components
in the x, y and z directions respectively, p is the impression
pressure and is the apparent shear viscosity. T is the
temperature while C
p
and k are the specic heat and thermal
conductivity of the material, respectively. The subscripts l and
s refer to liquid and solid phases, respectively.
It should be pointed out that in most ow analysis
programs for injection molding such as Moldow, the material
is assumed to be incompressible during the lling stage. This
assumption means that the density of the melt is constant and
thus the rst term in equation (1) is ignored. For convenience,
this type of ow is called the incompressible ow and the
ow of a polymer melt which obeys equation (1) is called
the compressible ow in this paper.
The interfacial energy balance equation for an amorphous
polymer can be written as [6]
k
l
T
l
z
= k
s
T
s
z
. (6)
For simplicity, we assume symmetry of the ow eld about
the mid plane in the thickness direction. This assumption can
easily be relaxed if necessary by making minor adjustment
2
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
of the appropriate equations. As such, for a given viscosity,
by integrating the momentum equations (2) and (3) along
the thickness direction as well as applying the no slip
conditions at the wall, the following equations can be
derived:
u =
x
, (7)
v =
y
, (8)
where
x
=
p
x
,
y
=
p
y
and =
_
h
z
z

d z.Substituting
the velocities into the continuity equation, the pressure
equation can be derived as [14]

x
_

S
p
x
_
+

y
_

S
p
y
_
G
p
t
= F, (9)
where the uidity

S, G and F are given by

S =
_
h
0
dz (10)
G =
_

0
_

l
p
_
T
dz +
_
h

s
p
_
T
dz (11)
F =
_

0
_

l
T
_
p
T
t
dz +
_
h

s
T
_
p
T
t
dz
+ (
l

s
)
z=

t
. (12)
Here, is the half-thickness of the molten layer.
2.2. Material models for polymer melt
The widely accepted Cross-WLF model, which is accurate
over a wide temperature range, is adopted to characterize
the non-Newtonian behavior of the molten polymer [14, 19,
20]:
=

0
(T, p)
1 + (
0
/

)
1n
, (13)
where

is the shear stress at the transition between Newtonian


and power-law behavior.
0
is the zero-shear-rate viscosity
with temperature and pressure dependent factors, which can
be expressed as

0
(T, p) = D
1
exp
_

A
1
[T T

(p)]
A
2
+ [T T

(p)]
_
, (14)
where
T

(p) = D
2
+ D
3
p (15)
A
2
(p) =

A
2
+ D
3
p. (16)
Here, n,

, D
1
, D
2
, D
3
, A
1
and

A
2
are seven material constants
to be determined from experimental measurements.
The density of the polymer melt is a function of both
temperature and pressure. For an amorphous polymer, this
could be described by the two-domain Tait equation of state
V(T, p) = V
0
(T )
_
1 C ln
_
1 +
p
B(T )
__
, (17)
Figure 2. Control volume and nite difference grid.
where C has a constant value of 0.0894. V
o
(T) and B(T) take
the forms
V
0
(T ) =
_
b
1,l
+ b
2,l
(T b
5
) if T T
t
(p)
b
1,s
+ b
2,s
(T b
5
) if T < T
t
(p)
(18)
B(T ) =
_
b
3,l
exp[b
4,l
(T b
5
)] if T T
t
(p)
b
3,s
exp[b
4,s
(T b
5
)] if T < T
t
(p)
(19)
T
t
(p) = b
5
+ b
6
p, (20)
where T
t
is the pressure-dependent glass transition temperature
for an amorphous polymer. b
1,l
, b
2,l
, b
3,l
, b
4,l
, b
1,s
, b
2,s
, b
3,s
,
b
4,s
, b
5
, b
6
are material constants.
2.3. Numerical implementation
Following the approach of Hieber and Shen [1] and Chiang
et al [14], a combined nite element/nite difference/
control volume scheme is employed. Specically, the
nite element technique is used to calculate the pressure in
the planar coordinates, the nite difference scheme is for the
discretizations of timewise and gapwise coordinates and the
control volume approach is for the advancement of the melt
front.
The impression geometry is meshed into thin triangular
elements. Each element is further divided into three sub-
domains by connecting the midpoints of its three edges to its
centroid. The control volume at each node is the union of all
sub-volumes that are linked to that node (see the hatched area
in gure 2).
The numerical equation for the net mass ow m
(l)
of
element l can be written as
m
(l)
i
=

S
(l)
3

i=1
D
(l)
ik
p
N
+
A
(l)
G
(l)
3
3

i=1
E
(l)
ik
p
N
p
N

t
+
A
(l)
F
(l)
3
, i = 1, 2 or 3
(21)
where

S
(l)
and D
(l)
ik
are the owconductance and the coefcient
of the nodal pressure to the net ow in element l, respectively.
3
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
A
(l)
is the area of the element l. Mass conservation for the
control volume at node N is given by [14]
l


S
(l)
3

i=1
D
(l)
ik
p
N
+
1
3
l

A
(l)
G
(l)
3

i=1
E
(l)
ik
p
N
p
N

t
+
1
3
l

A
(l)
F
(l)
= 0. (22)
Equation (22) is solved for the nodal pressure.
The energy equation is solved at each node to obtain the
temperature prole. This equation is discretized in both time
t and spatial dimension z using the nite difference method.
Time discretization is linked with the growing scheme used
for capturing the melt front while the spatial discretization is
shown in gure 2. Each element has the same number of
points for thermal calculations.
The nodal energy equation with the averaged temperature
evaluated at the node of each element can take the form as [5]
[K]
N
{M
k+1
}
N
= {F}
N
. (23)
Here [K]
N
, {M
k+1
}
N
and {F}
N
are the global stiffness matrix,
the pending matrix vector at time step k+1 and the force vector,
respectively. The force vector includes the convection, viscous
heating and the initial temperature terms. Equation (23)
can be solved for temperature.
A ll factor f
ij
which is assigned to every nodal control
volume to indicate the percentage of lling is dened by
f
ij
= V
ij
_
V
i
, (24)
where V
ij
and V
i
are the occupied volume and the control
volume at the ith node in the jth time step, respectively. f
ij
has
a value between 0 and 1. Based on the ll factor, all nodes are
classied as either completely lled node (f
ij
= 1), front node
(0 < f
ij
< 1) or empty node (f
ij
= 0). The node(s) through
which the melt enters the impression are dened as the gate
node(s) with f
ij
= 1.
In each time step, the pressure eld is solved to obtain
the velocity distribution in the ow domain. It is important to
note that for T(z) in the newly lled volume, its initial value
may be taken as the average value of its neighbor upper-stream
element.
3. Experimental procedure
A 25 ton Battenfeld injection molding machine HM25/60 S
UNILOG B6 was used for the experiments. The maximum
injection pressure, injection rate and injection speed of the
machine are 260 MPa, 47 cm
3
s
1
and 45 cms
1
, respectively.
A micro-thickness rectangular regular-sized part was chosen
to demonstrate the effects of compressibility during the lling
stage. The dimensions of the plate are 40 mm 24 mm
(see gure 3) with part thickness varying from 920 to
370 m. Six levels of thickness, namely, 920, 720, 620, 520,
420 and 370 m were investigated. It should be noted that
370 m was the smallest thickness that could be successfully
molded by the employed machine and processing conditions
chosen.
Figure 3. Part geometry and location of the pressuretemperature
sensor.
Figure 4. Layout of the core insert.
Table 1. Properties of COC TOPAS 5013L-10.

l

s
T
g
C
p
K
(g cm
3
) (g cm
3
) (

C) (J kg
1
K
1
) (W m
1
K
1
)
0.9044 1.0126 125 2777.667 0.194
A two-plate mold was fabricated to produce the designed
plastic part. To adjust the part thickness, the same mechanism
as described by Ong et al [11] was used. As shown in
gure 4, carbon steel shims with different thicknesses, such as
50, 100, 200, 500 m, etc were placed between the primary
insert and secondary insert to control the thickness of the
impression. Adding more shims decreased the impression
thickness and thus the part thickness. The polymer melt
was injected into the mold impression through the side gate.
One pressuretemperature sensor (Kistler Type 6190CA) was
mounted near the gate to measure the impression pressure and
mold wall temperature during the lling phase (see gure 3).
A data acquisition system (Kistler CoMo-Injection) was also
employed to obtain the pressure and temperature data.
Due to its superior properties such as relatively high
glass transition temperature, high transparency, low water
absorption, etc, the TOPAS family has recently received
increasing attention from both academic and industry as a
promising material for microuidic applications [21]. Thus
high-ow COC TOPAS 5013L-10 produced by TOPAS
Advanced Polymers was chosen as a model material in this
study (see table 1 for its main properties). The processing
conditions for different thicknesses are tabulated in table 2.
The mold temperature (46

C), melt temperature (300

C),
injection speed (20 cm s
1
) and cooling times (15 s) were set
the same in all cases. For each change in part thickness, the
machine was cycled until the whole process was stable. Then
the parts of ve consecutive runs were collected to ensure the
repeatability of the measurements.
4
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
Figure 5. Finite element mesh of impression geometry.
Table 2. Processing conditions for different part thicknesses.
Thickness Injection pressure Injection rate
(m) (MPa) (cm
3
s
1
)
920 40 20
720 70 10
620 80 15
520 80 10
420 140 20
370 150 30
Table 3. Tait PVT constants for COC TOPAS 5013L-10.
Symbol Quantity
b
1,l
(m
3
kg
1
) 0.001 012
b
2,l
(m
3
kg
1
K
1
) 6.693 000 10
7
b
3,l
(Pa) 1.690 210 10
+8
b
4,l
(1/K) 0.004 276
b
1,s
(m
3
kg
1
) 0.001 012
b
2,s
(m
3
kg
1
K
1
) 2.310 000 10
7
b
3,s
(Pa) 2.805 470 10
+8
b
4,s
(1/K) 0.002 380
b
5
(K) 402.55
b
6
(K/Pa) 4.694 000 10
7
4. Simulation inputs
Figure 5 shows the nite-element mesh of the impression
which has 1280 regular triangular elements and 693 nodes. As
the pressure transducer measured the average pressure over an
area of 4 mm diameter, the corresponding average pressure in
the simulation was calculated by taking the average value of
the seven nodes within the sensor area (see the seven small
dots in gure 5).
A reliable numerical analysis can only be conducted with
accurate material data. Generally, in most simulations, the
effects of pressure on the material rheological properties are
normally neglected. As relatively high injection pressure
was used in our experiments, the effect of pressure-dependent
viscosity on impression pressure may become important [22].
Thus, material characterization was conducted by Autodesk
Australia Pty. Ltd to obtain the thermal properties, PVT
data and especially the pressure-dependent viscosity of COC
TOPAS 5013L-10 (see tables 3, 4, and gures 6 and 7). As the
Figure 6. PVT data for COC TOPAS 5013L-10.
injection time is short (0.150.25 s), specic heat and thermal
conductivity of the mold material and polymer melt, the mold
temperature as well as heat transfer coefcient at the interface
between the melt and the mold are assumed to be constant
during the lling phase. The glass transition temperature was
chosen as the no-ow temperature.
The density, thermal conductivity and specic heat of
the mold material are 7820 kg m
3
, 32 W m
1
K
1
and
500 J kg
1
K
1
, respectively.
5. Results and discussions
For convenience, the time when the pressure signal started
to increase was set to zero. As shown in gure 8, for each
level of thickness, three measured pressure traces (curves)
during the lling stage are compared with predicted impression
pressures for both incompressible ow (unlled triangle) and
compressible ow (unlled circle). It should be noted that
different chart scales are used for the last two thicknesses, i.e.
420 and 370 m, and that the measured pressure traces in
gure 8(f ) are expanded to the rst second of the molding
cycle. Figure 9 presents the bulk temperature, pressure
and density variations with lling time within the pressure
transducer area for different part thicknesses.
For thickness larger than 620 m (920, 720 and
620 msee gures 8(a), (b) and (c), respectively), the
numerical results are in good agreement with the experimental
measurements. There is little pressure difference predicted
between incompressible ow and compressible ow. Figure 9
also indicates that there is little variation in bulk density with
lling time (0.8920.916 g cm
3
) within the sensor area for
these thicknesses. This is understandable as the corresponding
bulk temperature is almost constant (299.1298.4

C) and
the pressure variation is small (027 MPa). These results
conrm that the effects of compressibility can be ignored in
conventional macro-injection mold lling.
Table 4. Cross-WLF constants for COC TOPAS 5013L-10.
n

(Pa) D
1
(Pa.s) D
2
(K) D
3
(K/Pa) A
1

A
2
(K)
0.402 71 46 129.8 4.511 08 10
+17
343.150 1.200 10
7
44.743 51.600
5
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
(a) (b)
Figure 7.Viscosity of COC TOPAS 5013L-10 as a function of shear rate and (a) temperature, (b) pressure.
(a) (b)
(c) (d)
(e) ( f )
Figure 8. Comparison of simulated (symbols) and experimental (curves) impression pressure for part thicknesses (a) 920 m, (b) 720 m,
(c) 620 m, (d) 520 m, (e) 420 m and (f ) 370 m.
6
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
Figure 9. Predicted variation of bulk temperature, pressure and density with lling time within the pressure transducer area for different part
thicknesses.
For the last three thicknesses (520, 420 and 370 msee
gures 8(d), (e) and (f ), respectively), although simulated
results with the compressible ow or incompressible ow
agree with experimental results rather well at initial lling,
it can be observed that as lling progresses, the compressible-
ow-simulated results t the measured pressure traces much
better than the incompressible-ow-simulated results. It
should be highlighted here that at the very end of the lling
phase, there is always a pressure surge measured and predicted
when the impression is completely lled (see gure 8(f )). As
this occurs over a relatively short time duration, the exact value
measured and predicted over time might not be very precise.
These major differences in lling behavior for thin
part predicted between compressible and incompressible
predictions are because the polymer melt tends to solidify
more easily in thinner parts due to higher thermal loss through
the mold wall. As a result, much higher lling pressure is
required, especially in late lling, for the closely packed melt
to ow in the thin impression. This will result in more obvious
variation in bulk density with a decrease in thickness due to this
much higher injection pressure effect. As shown in gure 9,
although the bulk temperatures within the pressure transducer
area for thicknesses 520, 420 and 370 m generally drop
slightly (298.7295.1

C), there is a trend that the temperature


reduces considerably in late lling for part thickness of
370 m. In addition, it is obvious that the impression
pressure increases signicantly as the part thickness decreases,
i.e. 036.2, 068.1 and 095.6 MPa for part thicknesses
7
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
(a) (b) (c)
Figure 10. Predicted bulk distributions within the molded part after the lling phase for part thickness of 370 m for (a) temperature,
(b) pressure, and (c) density.
520, 420 and 370 m, respectively. The corresponding bulk
density variation is therefore more pronounced, e.g., 0.897
0.963 g cm
3
for part thickness of 370 m. Figures 10(a),
(b) and (c) show the predicted bulk temperature, pressure and
density distributions, respectively, within the impression after
the lling phase for the compressible ow of part thickness
370 m. It should be noted that the density distribution is
dependent on the proximity of the location to the gate. As
the area near the gate experiences higher pressure, the effect
of compressibility in this region is more pronounced and thus
the packing density is higher (see the bottom-center area in
gure 10(c)).
As such, ignoring the effects of melt compressibility may
lead to considerable inaccuracy in the simulation of the lling
stage in micro-injection molding, for both pressure and density
predictions.
6. Conclusions
This study presents both numerical simulation and
experimental investigation on the compressible ow of COC
TOPAS 5013L-10 polymer in the mold lling of micro-
thickness impression. Our investigation indicates that the
assumption of the incompressible lling ow is sufciently
accurate for macro-injection molding, but not for micro-
injection molding. The compressibility of the molten polymer
should be considered in micro-injection mold lling as it
has important effects on impression pressure and density
distribution in the lled part. The effect of melt compressibility
becomes more pronounced with a decrease in part thickness
and that the density distribution in the impression depends on
the relative distance to the gatethe closer the location to the
gate, the higher the compression and melt density. Further
investigations on other type of amorphous polymers as well
as semi-crystalline polymers will be conducted to evaluate the
pressure (and thus compressibility) sensitivity of the various
polymers.
Acknowledgments
The authors gratefully acknowledge the research funding
support from the Singapore-MIT Alliance Program and for
material testing conducted by Autodesk Australia Pty. Ltd.
Thanks are also extended to the Manufacturing Process Lab
1 & 2 (NTU) for assistance in manufacturing the mold insert
and for carrying out the various experiments.
References
[1] Hieber C A and Shen S F 1980 A
nite-element/nite-difference simulation of the
injection-molding lling process J. Non-Newton. Fluid
Mech. 7 132
[2] Tadmor Z, Broyer E and Gutnger C 1974 Flow analysis
network (FAN)a method for solving ow problems in
polymer processing Polym. Eng. Sci. 14 6605
[3] Wang V W, Hieber C A and Wang K K 1986 Dynamic
simulation and graphics for the injection molding of
three-dimensional thin parts J. Polym. Eng. 7 2145
[4] Isayev A I 1987 Injection and Compression Molding
Fundamentals (New York: Marcel Dekker)
[5] Kennedy P 1995 Flow Analysis of Injection Molds (New York:
Hanser)
[6] Han K H and Im Y T 1997 Compressible ow analysis of
lling and post-lling in injection molding with
phase-change effect Compos. Struct. 38 17990
[7] Lam Y C, Chen X, Tam K C and Yu S C M 2003 Simulation of
particle migration of powder-resin system in injection
molding J. Manuf. Sci. Eng. 125 53847
[8] Yu L, Koh C G, Lee L J, Koelling K W and Madou M J 2002
Experimental investigation and numerical simulation of
injection molding with micro-features Polym. Eng. Sci.
42 87188
[9] Yu L,, Lee L J and Koelling K W 2004 Flow and heat transfer
simulation of injection molding with microstructures
Polym. Eng. Sci. 44 186676
[10] Wu C H and Wu S M 2007 Flow analysis of micro-channel in
injection molding J. Polym. Eng. 27 10727
[11] Ong N S and Koh Y H 2005 Experimental investigation into
micro injection molding of plastic parts Mater. Manuf.
Process. 20 24553
8
J. Micromech. Microeng. 21 (2011) 095019 Q M P Nguyen et al
[12] Giboz J, Copponnex T and M el e P 2007 Microinjection
molding of thermoplastic polymers: A review
J. Micromech. Microeng. 17 96109
[13] Attia U M, Marson S and Alcock J R 2009 Micro-injection
moulding of polymer microuidic devices Microuid.
Nanouid. 7 128
[14] Chiang H H, Hieber C A and Wang K K 1991 A unied
simulation of the lling and postlling stages in injection
molding: part 1. Formulation Polym. Eng. Sci.
31 11624
[15] Chiang H H, Hieber C A and Wang K K 1991 A unied
simulation of the lling and postlling stages in injection
molding: part II. Experimental verication Polym. Eng. Sci.
31 12539
[16] Changyu S, Lixia W and Qian L 2007 Numerical simulation of
compressible ow with phase change of lling stage in
injection molding J. Reinf. Plast. Compos. 26 35372
[17] Tham N C, Juttner G, Loser C, Pham T and Gehde M 2010
Determination of the heat transfer coefcient from
short-shots studies and precise simulation of microinjection
molding Polym. Eng. Sci. 50 16573
[18] Chen B S and Liu W H 1994 Numerical simulation of the
post-lling stage in injection molding with a two-phase
model Polym. Eng. Sci. 34 83546
[19] Cross M M 1965 Rheology of non-Newtonian uids: a new
ow equation for pseudoplastic systems J. Colloid. Sci.
20 41737
[20] Williams M L, Landel R F and Ferry J D 1955 The
temperature dependence of relaxation mechanisms in
amorphous polymers and other glass-forming liquids J. Am.
Chem. Soc. 77 37017
[21] Blochowiak M, Pakula T, Butt H J, Bruch M and Floudas G
2006 Thermodynamics and rheology of cycloolen
copolymers J. Chem. Phys. 124 13
[22] Fern andez M, Munoz M E, Santamaria A, Syrjala S and Aho J
2009 Determining the pressure dependency of the viscosity
using PVT data: a practical alternative for thermoplastics
Polym. Test. 28 10913
9

S-ar putea să vă placă și