Sei sulla pagina 1di 9

Determination of the Heat Transfer Coefcient From

Short-Shots Studies and Precise Simulation of


Microinjection Molding

Tham Nguyen-Chung,1 Gabor Juttner,2 Cindy Loser,1 Tung Pham,3 Michael Gehde1
1
Chemnitz University of Technology, D-09107 Chemnitz, Germany
2

Kunststoff-Zentrum in Leipzig, Erich-Zeigner-Alle 44, D-04229, Leipzig, Germany

Borealis Polyolene GmbH, St.-Peter-Str. 25, A-4021 Linz, Austria

The lling process of a micro-cavity was analyzed by


modeling the compressible lling stage by using pressure-dependent viscosity and adjusted heat transfer
coefcients. Experimental lling studies were carried
out at the same time on an accurately controlled
microinjection molding machine. On the basis of the
relationship between the injection pressure and the lling degree, essential factors for the quality of the simulation can be identied. It can be shown that the ow
behavior of the melt in a micro-cavity with a high aspect ratio is extremely dependent on the melt compressibility in the injection cylinder. This phenomenon
needs to be considered in the simulation to predict an
accurate ow rate. The heat transfer coefcient
between the melt and the mold wall that was determined by the reverse engineering varies signicantly
even during the lling stage. With increasing injection
speed and increasing cavity thickness, the heat transfer coefcient decreases. It is believed that the level of
the cavity pressure is responsible for the resulting heat
transfer between the polymer and the mold. A pressure-dependent model for the heat transfer coefcient
would be able to signicantly improve the quality of
the process simulation. POLYM. ENG. SCI., 50:165173,
2010. 2009 Society of Plastics Engineers

INTRODUCTION
Microinjection molding is nowadays seen as the most
efcient technology for large-scale manufacturing of therPresented in part at the XVth International Congress on Rheology, 2008,
Monterey, California.
Dedicated to Prof. Dr.-Ing. G. Mennig on the occasion of his 70th birthday.
Correspondence to: Tham Nguyen-Chung; e-mail: tham.nguyen.chung@
mb.tu-chemnitz.de
DOI 10.1002/pen.21536
Published online in Wiley InterScience (www.interscience.wiley.com).
C 2009 Society of Plastics Engineers
V

POLYMER ENGINEERING AND SCIENCE-2010

moplastic polymer micro-parts. The rapid development of


such micro-engineering technologies requires new concepts for molding machines and micro-molds on the one
hand and a better understanding of the process characteristics on the other. Molding micro-features of high aspect
ratios still remains a challenge in particular as a result of
the high surface-to-volume ratio of the part that makes
the melt freeze-in much faster than in conventional
injection molding.
The lling and solidication of macro-scale injection
moldings can be predicted by using commercial computer-aided engineering software. It is also known that the
conventional tools do not work well for all process conditions as applicable for microinjection molding. The reasons for the discrepancies between simulation and experimental results might be the lack of a high quality database used in the simulation and the improperly specied
boundary conditions that do not reect the real state in
the cavity [14].
In this article, aspects related to the boundary conditions, especially the thermal contact behavior, were taken
into consideration. It is known for macro-scale injection
molding that the heat transfer coefcient at the interface
between polymer and the mold changes with time during
an injection cycle. Typical values for the lling stage
range between 5000 and 2000 W/m2K [5, 6]. Smaller values were also reported by others [7, 8].
Some authors [9, 10] calculated the thermal contact resistance between the polymer and the mold by solving the
one-dimensional heat transfer equation by using the measured mold surface temperature as the thermal boundary
condition. The values of the heat transfer coefcient calculated this way related to the whole injection molding
process including the packing and cooling stage. It is not
always possible to use the same heat transfer coefcient
for different molding stages.

Especially during the packing and cooling stage, an air


gap partial between the polymer melt and the mold wall
might arise as a result of the thermal shrinkage, leading
to a drop of the heat transfer coefcient [911]. This
could even happen earlier in a micro-cavity as a result of
the fast cooling [12]. On the other hand, it is also possible
that the high injection pressure that is needed to overcome
the high ow resistance in the micro-cavity might prevent
an earlier shrinkage in the micro-cavity.
Rhee et al. [13] reported that the heat transfer coefcient can increase signicantly from 11,700 W/m2K to
28,500 W/m2K when a mold-releasing agent is used as an
interstitial medium between the solid polymer and two
steel blocks. This phenomenon was explained by the
inuence of the contact condition on the heat transfer
condition. Zhang et al. [14] have also found that the
roughness of the mold surface has a signicant effect on
the volume of the micro-cavity and the heat transfer
between the polymer melt and the mold.
Yao and Kim [15] investigated the inuence of the
micro-scale phenomena, specically size-dependent viscosity, wall slip, and surface tension, on the lling process of polymer melts in a micro-cavity. The inuence of
the crystallization on the thermal contact resistance was
reported by Golf et al. [16]. Other works [17, 18] suggested that the thermal conductivity is enhanced along the
ow direction and diminished across the ow. As indicated by Dai and Tanner [19], only small effects were
found for sheared polypropylene.
This article aims to clarify the inuence of the boundary conditions on the results of the microinjection molding
simulation by using commercially available software. The
heat transfer coefcient in the lling stage was identied
by means of reverse engineering based on the relationship
between the injection pressure and the lling degree. This
could be provided by short-shot studies conducted with a
precisely controlled microinjection molding machine.
Another important issue regarding the boundary condition that has not been considered properly in previous
works is the melt compression caused by the high pressure
level in the injection barrel. It might not only increase the
temperature of the melt in the barrel as a result of the
compression heating but also reduce the actual volume
rate in the cavity, even if a constant plunger speed is
applied. Consequently, this inuences the lling behavior
that is especially critical for micro-parts of high aspect
ratios. The melt compression in the injection barrel needs
to be included in the simulation to properly investigate the
behavior of the heat transfer in the cavity.

EXPERIMENTS
Material Data
Two grades of isotactic polypropylene (HC600TF and
DM55 pharm) with different viscosities and crystallinity
166 POLYMER ENGINEERING AND SCIENCE-2010

degrees have been provided by Borealis Polyolene


GmbH (Linz, Austria). Material characterization was conducted by Borealis, Moldow Pty. (Kilsyth, Australia)
and the Lehrstuhl fur Kunststofftechnik at the FriedrichAlexander University Erlangen-Nurnberg (Erlangen, Germany). Besides standard testing procedures to obtain
pvT-data and the thermal properties, a special test was
conducted by Moldow to determine the viscosity at high
shear rates (up to 1.6 3 106 s21) as well as the pressure
dependency of the viscosity. The material data obtained
for HC600TF have been published [20], and they have
also been included into the database of the Moldow software since the version 6.2 of Moldow Plastics Insight.
Therefore, only data for DM55 pharm will be shown in
the following.
The viscosity g was approximated as a function of the
temperature T, the shear rate c_ , and the pressure p by
using the Cross-WLF-model [21]:
Z

Z0
1

Z0 1n

t g

(1)



A1 T  T 
;
Z0 D1 exp 
A2 T  T 

(2)

T  D2 D3 p;

(3)

A2 A~2 D3 p;

(4)

where n, t*, D1, D2, D3, A1, and A~2 are data-tted coefcients: n 0.2336, s* 36932.7 Pa, D1 1.1233 3 1017
Pa s, D2 263.15 K, D3 1.5 3 1027 K/Pa, A1
37.761, and A~2 5 51:6 K.
The pvT data were approximated by using the Tait
model [21]:



p
ut T; p; (5)
uT; p u0 T 1  C ln 1
BT
where u(T, p) is the specic volume at temperature T and
pressure p, u0(T) is the specic volume at zero gauge
pressure, C is a constant (C 0.0894), and B accounts
for the pressure sensitivity of the material and as dened
below.
The upper temperature region (T Tt) can be described
by the equations:
u0 T b1m b2m T  b5 ;

(6)

BT b3m expb4m T  b5 ;

(7)

ut T; p 0;

(8)

where Tt is the transition temperature, b1m, b2m, b3m, b4m,


and b5 are data-tted coefcients. The coefcient b5 represents the volumetric transition temperature at zero
gauge pressure.
DOI 10.1002/pen

TABLE 1. Thermal conductivity of polypropylene DM55 pharm.


Temperature (8C)

Thermal conductivity (W/mK)

279.2
258.9
238.7
197.6
177.6
157.7
138.5
118.3
98.3
78.2
58.5
38.8

0.1697
0.1676
0.1639
0.1547
0.1556
0.1552
0.2244
0.2267
0.2329
0.2371
0.2368
0.2442

FIG. 1. Micro-spiral.

The lower temperature region (T \ Tt) can be


described by the equations:
u0 T b1s b2s T  b5 ;

(9)

BT b3s expb4s T  b5 ;

(10)

ut T; p b7 expb8 T  b5  b9 p;

(11)

where b1s, b2s, b3s, b4s, b5, b7, b8, and b9 are data-tted
coefcients.
The transition temperature Tt is a linear function of the
pressure:
Tt p b5 b6 p:

(12)

The values of the above mentioned coefcients are as


follows: b1m 0.001311 m3/kg, b2m 1.2131026 m3/
kgK, b3m 5.757313107 Pa, b4m 0.005009 1/K, b1s
0.001228 m3/kg, b2s 7.79731027 m3/kgK, b3s
6.708583107 Pa, b4s 0.004396 1/K, b5 393.15 K,
b6 2.631027 K/Pa, b7 8.29331025 m3/kg, b8
0.2045 1/ K and b9 5.4713101028 1/Pa.
The thermal conductivity and the specic heat are provided as discrete values at a wide range of temperatures
(Tables 1 and 2).

Part Geometries
Two different geometries were used: the micro-spiral
(Fig. 1) and the micro-plate (Fig. 2). The micro-spiral has
a ow length of 32 mm and a width of 1.5 mm. Two different thicknesses (0.2 mm and 0.5 mm) are available.
The micro-plate is a rectangular geometry of 12 mm 3
11 mm 3 0.5 mm. For both dimensions, the cavity is fed
from a conical sprue. The sprue and the part cavity are
connected through a junction zone.

Molding
Molding operations were conducted with a plunger
injection molding machine (formicaPlast) developed by
the Kunststoff-Zentrum in Leipzig (Germany) [22]. The
injection unit consists of a pre-plasticizing plunger and an
injection plunger (Fig. 3). The diameter of the injection

TABLE 2. Specic heat of polypropylene DM55 pharm (cooling rate,


0.33 K/s).
Temperature (8C)

Specic heat (J/kgK)

280
192
125
118
114
111
108
105
102
98
89
60
32

3030
2851
2572
2897
4349
7709
17,814
9404
3441
2661
2390
2033
1762

DOI 10.1002/pen

FIG. 2. Micro-plate.

POLYMER ENGINEERING AND SCIENCE-2010 167

for every ve short-shots with the same conditions. The


lling degree was determined by the surface area of the
scanned images of the moldings related to the corresponding surface area of a full part. A correction was
needed to take into account the sprue and the junction
region.

SIMULATION
Model Geometries

FIG. 3. Schematic of the plunger injection unit.

plunger is 3 mm, and a fast electrical drive is used, ensuring a high precision of control for the injection speed and
the plunger position. The maximum injection pressure and
injection rate of the machine are 300 MPa and 3.5 cm3/s,
respectively.
Short-shots of different lling degrees (Fig. 4) were
produced by adopting different stop positions for the
injection plunger. Five shots were made for each set of
the processing parameters. The melt and mold temperatures as well as the plunger speed in the short-shot experiments were varied as shown in Table 3. No packing pressure was applied. The cavity pressure was measured by
using a miniaturized quartz sensor (KISTLER 6183AE).
The pressure sensor was located at the junction zone
between the sprue and the part cavity (Fig. 5) for both
part geometries.
Figure 6 shows the cavity pressure for short-shots with
10 different stop positions of the plunger. For every
plunger stop position ve short-shots were conducted,
resulting in a total number of 50 for the pressure curves.
The corresponding pressure curves of the same plunger
stop position are almost identical except for the last segments of the curves, which indicated a high degree of
reproducibility of the experiments. Average values of the
maximum pressure and the lling degree were obtained

Three-dimensional lling simulations for the micro-spiral and the micro-plate were conducted by using the software Moldow Plastics Insight (version 6.1; revision 3
was used for the micro-spiral and version 6.2, revision 2
for the micro-plate).
Only half of the actual cavity was taken into account
(Fig. 7) in cases of the micro-spiral geometry. The symmetry surface was considered as a usual mold wall with a
no-slip condition applied being an acceptable simplication as a result of the fact that the dimensions of the
junction zone are much larger compared with the thickness of the spiral region. The conical sprue was not
included in the model because the ow resistance in the
sprue is much smaller than the one in the spiral and
therefore can be neglected. This allows the calculated
injection pressure to be related to the cavity pressure
measured at the junction zone between the sprue and the
micro-spiral (Fig. 5).
In case of the micro-plate, the whole cavity including
the sprue was considered (Fig. 8) because the pressure
loss in the sprue cannot be negligible as compared with
the pressure loss in the micro-plate. The pressure level in
the micro-plate is much lower than in the micro-spirals,
as can be seen later in the results.

Governing Equations
The governing equations used in the software are the
conservation equations of mass, momentum, and energy
[23]:
qr
r  r~
u 0
(13)
qt
Dr~
u
rp r  t
Dt

(14)

FIG. 4. Short-shots of the micro-plate.


168 POLYMER ENGINEERING AND SCIENCE-2010

DOI 10.1002/pen

TABLE 3. Processing conditions.

No.

Material

Geometry

d
(mm)

Tmold
(8C)

Tmelt
(8C)

vP
(mm/s)

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16

HC600TF

Micro-spiral

0.2

20/30a

280

20
50
100
200
20
50
100
200
50
100
150
200
50
100
150
200

DM55 pharm

Micro-plate

0.5

20/30a

280

0.5

70

220

280

d, thickness; Tmold, mold temperature; Tmelt, melt temperature;


vP, plunger speed.
a
The mold temperatures are different at two halves of the mold (higher
temperature at the nozzle side to avoid melt-freezing at the nozzle).

rcp

DT
Dp
bT
r  krT Zg2
Dt
Dt

(15)

where t, ~
t, T, p, s; c_ , q, g, cp, b, and k denote time, velocity vector, temperature, pressure, stress tensor, shear rate,
density, viscosity, specic heat, expansivity, and thermal
conductivity, respectively.
For the constitutive equation, the generalized Newtonian uid model is assumed [23]:
t 2Zg ; T; pg

(16)

where the rate of strain tensor c_ is dened as follows:


g


1
r~
u r~
uT
2

FIG. 6. Cavity pressure measured during short-shot studies (10 lling


degrees, 5 shots per lling degree).

Heat Transfer Coefcient


For each processing condition, simulations with different heat transfer coefcients were carried out, providing
the relationships between the cavity pressure and the lling degree. These were then compared with the corresponding experimental data. Figure 9 shows an example:
The results calculated with three different heat transfer
coefcients indicate that using a heat transfer coefcient
between 0 and 1500 W/m2K would lead to a more precise
prediction of the pressure than using the default value of
5000 W/m2K.
Geometry and Time Discretization
The geometry and time discretization should be
adjusted properly to capture the highly changing velocities and temperatures during the lling process. For
example, in case of the micro-spiral the average mesh

(17)

Melt Compression in the Injection Cylinder


To precisely consider the compressibility of the melt in
the injection barrel, information about the plunger diameter, its starting position, and its velocity should be provided (Fig. 3). The actual melt temperature and the actual
ow rate of the melt owing from the injection barrel into
the cavity were calculated by the software at every time
step on the basis of the pressure at the injection location.

FIG. 5. Location of the pressure sensor.

DOI 10.1002/pen

FIG. 7. Model for the micro-spiral.

POLYMER ENGINEERING AND SCIENCE-2010 169

FIG. 8. Model for the micro-plate.

size on the surface for the part was set to 0.1 mm, and at
least 12 layers of tetrahedra were needed along the spiral
thickness. The mesh of the micro-spiral part consisted of
about 500,000 tetrahedral elements in total. Similarly, the
maximal increase of the volume between two successive
time steps was reduced from the default value of 4% to
0.025% of the cavity volume.
RESULTS
Figure 10 shows the relationship between the injection
pressure and the lling degree for the micro-spirals with a
thickness of 0.2 mm. The experiments were done with a
plunger speed varying between 20 mm/s and 200 mm/s,
while all other processing parameters were kept constant.
For injection speeds between 50 mm/s and 200 mm/s, the
experimental data can be properly tted by the simulation
by using values of the heat transfer coefcient between
25,000 W/m2K and 1500 W/m2K.
The heat transfer coefcient value of 25,000 W/m2K
seems, at rst glance, to be very high, but not unrealistic.
In the older versions of the Moldow software, this value

FIG. 9. Determining the heat transfer coefcient by means of reverse


engineering.

170 POLYMER ENGINEERING AND SCIENCE-2010

FIG. 10. Pressure vs. lling degree for the micro-spirals with a thickness of 0.2 mm.

was used as the default value corresponding to a perfect


contact condition between the molten polymer and the
mold. An even higher value for the heat transfer coefcient (28,500 W/m2K) was found by Rhee et al. [13] as
they used a mold-releasing agent to ll up the gaps
between the polymer melt and the mold rough surface. In
the present case, a good contact condition is obviously
created at an injection speed of 50 mm/s.
It can be seen that the determined heat transfer coefcient systematically decreases with increasing injection
speed. Because the position of the curves clearly shows
that a higher cavity pressure corresponds to a higher heat
transfer coefcient, it is believed that the pressure level in
the cavity is responsible for the thermal contact behavior
between the polymer and the mold.
An exception is the case with the lowest injection
speed of 20 mm/s where the experimental data cannot be
met by using any value of the heat transfer coefcient.
With a heat transfer coefcient of 14,000 W/m2K, the
maximum lling degree would be predicted correctly, but

FIG. 11. Pressure vs. lling degree for the micro-spirals of 0.2 mm
thickness for the lowest injection speed.

DOI 10.1002/pen

FIG. 12. Pressure distribution in the middle layer of the micro-spiral at


three different times of the lling process.

the calculated injection pressures at other lling degrees


were lower than the measured ones (Fig. 11).
The reason for that might be the strong variation of the
pressure and the frozen layer in the cavity. Figures 12
and 13 show the pressure and temperature distribution at
three different times during the lling process as calculated with a constant heat transfer coefcient h 14,000
W/m2K. At an earlier stage of the lling process (Fig.
12a), the cavity pressure is at a lower level so that the
thermal contact between the polymer and the mold is not
established, consequently giving a lower expected heat
transfer coefcient. As soon as the cavity pressure
increases (Fig. 12b and c), a better thermal contact will
be established, leading to a higher heat transfer coefcient
near the injection location. The frozen-in area will
increase very fast, as indicated by the temperature distribution in Fig. 13b and c. Subsequently, the cavity pressure is no longer able to compensate for the shrinkage at
the outer surface or at the region far away from the injection location. The heat transfer coefcient is expected to
decrease.
In that case a uniform heat transfer coefcient would
be incapable of reecting the complex behavior of the
polymer at the mold wall, which is also inuenced by
many other factors like the surface roughness, the viscosities, and the shrinkage behavior of the polymer melt. It
can be assumed as a rst simplication that the experimental data will be better tted if the heat transfer coef-

cient is modeled as a function of the local pressure in the


cavity. This is particularly essential if the pressure in the
cavity varies strongly during the lling process.
The same tendency regarding the pressure dependency
of the heat transfer coefcients can be seen for the microspirals with a thickness of 0.5 mm (Fig. 14). The values
of the heat transfer coefcient found here are lower than
those in case of the thinner micro-spirals. The reason for
that is again the cavity pressure that is lower for the
thicker spirals, which leads to lower heat transfer coefcients.
In case of the micro-plates, the plunger speeds between
50 mm/s and 200 mm/s were applied. The mold temperature was kept constant at 708C, and the melt temperature
was 2808C. The relationship between the pressure and the
lling degree in Fig. 15 shows that the corresponding heat
transfer coefcient decreases from 6000 W/m2K to 4000
W/m2K when the plunger speed increases from 50 mm/s
to 200 mm/s. The pressure level is lower as compared

FIG. 13. Temperature distribution in the middle layer of the micro-spiral at three different times of the lling process.

FIG. 15. Pressure vs. lling degree for the micro-plates, Tmelt
2808C.

DOI 10.1002/pen

FIG. 14. Pressure vs. lling degree for the micro-spirals with a thickness of 0.5 mm.

POLYMER ENGINEERING AND SCIENCE-2010 171

FIG. 16. Pressure vs. lling degree for the micro-plates, Tmelt
2208C.

FIG. 18. Volume rate vs. lling degree for the micro-spirals with a
thickness of 0.2 mm.

with the one in the micro-spiral cavity and gives less variation of the heat transfer coefcient.
If a melt temperature of 2208C instead of 2808C was
applied for the micro-plates, the heat transfer coefcient
took values between 5000 W/m2K and 1500 W/m2K (Fig.
16). The heat transfer coefcient decreased, although the
pressure level in the cavity was slightly higher than in
case of the higher melt temperature of 2808C. This can be
explained by the higher viscosity of the melt. The higher
viscosity requires higher pressure to overcome the roughness at the mold surface (Fig. 17) to establish the same
level of thermal contact as compared with the previous
case. The roughness of the mold surface is also the
reason for the pressure dependency of the heat transfer
coefcient.
Figure 18 shows the volume rates calculated for the
micro-spirals with a thickness of 0.2 mm when different
plunger speeds were applied. As the melt lled the sprue
and the junction zone, the volume rates at the injection
location were equal to the theoretical values, resulting
from the plunger speed and the cross section of the

plunger. When the melt reached the spiral, the volume


rate decreased by about 50%, which is a result of the melt
compression in the injection barrel caused by the high
ow resistance in the cavity. Because the heat transfer
coefcient is strongly inuenced by the injection speed as
previously shown, neglecting the compression of the melt
in the injection barrel would lead to incorrect results
when investigating the heat transfer behavior in a microchannel. Consequently, special attention has to be paid to
this issue before conclusions about the inuence of other
factors on the heat transfer coefcient can be made.
Those factors are the viscoelastic behavior of the polymeric material and the anisotropic nature of the thermal
conductivity or the wall slip, whereby the last one has
not been observed within the range of the investigated
processing conditions.

FIG. 17. Contact reduction due to the surface roughness of the mold.

172 POLYMER ENGINEERING AND SCIENCE-2010

CONCLUSIONS
It can be shown that the heat transfer coefcient
between the polymer and the mold wall has a signicant
inuence on the simulation results. By using precise material data and considering the melt compression in the
barrel, the actual volume rate and the temperature of the
melt at the entrance of the cavity can be correctly calculated. It was able to determine the heat transfer coefcient
by means of reverse engineering based on the relationship
between the cavity pressure and the lling degree. The
heat transfer coefcient increases by decreasing cavity
thickness or injection speed. It is believed that the pressure level in the cavity is mostly responsible for the thermal contact between the polymer and the mold wall. A
pressure-dependent model for the heat transfer coefcient
would be more suitable to describe the thermal contact
behavior in microinjection molding, especially in case of
micro-cavities of high aspect ratio.
To take this phenomenon into consideration in the
numerical simulation, three different aspects have to be
considered: surface roughness of the mold, material propDOI 10.1002/pen

erties of the polymer in the molten and solidifying state,


as well as the pressure distribution along the mold wall.
As found by Zhang et al. [14], the surface roughness
depends signicantly on the equivalent height of the
roughness layer. It is less dependent on the specic
roughness prole. We plan to develop a model to describe
the equivalent height and the equivalent prole of the
roughness layer as a function of the roughness degree. On
the basis of the equivalent prole, the maximum contact
area can be analytically calculated by the local pressure
and the melt viscosity at the mold wall. The heat transfer
coefcient can be determined by the ratio between the
contact area and the corresponding area of the roughness
prole. When this ratio is equal to 1, the maximum possible value of the heat transfer coefcient is reached. Also
the stiffness and the shrinkage of the solidifying layer
need to be taken into account. The mold lling simulation
has to be coupled with a structural simulation to determine the normal pressure exerted on the mold wall. The
relation between the current pressure and the highest pressure that occurred in the molten stage can be seen as a
measure of the heat transfer condition for the increasing
frozen-in layer. Further work is going on to prove the feasibility of these ideas.
ACKNOWLEDGMENTS
The authors acknowledge the nancial support of the
Deutsche Forschungsgemeinschaft (Bonn, Germany)
within the framework of the project NG 64/3-1. We wish
to thank the Lehrstuhl fur Kunststofftechnik at the Friedrich-Alexander University Erlangen-Nurnberg (Erlangen,
Germany), Borealis Polyolene GmbH (Linz, Austria),
and Moldow Pty (Kilsyth, Australia) for conducting the
material testing. The rst author acknowledges discussions with Dr. F. Costa and Dr. R. Zheng on the numerical simulations with Moldow software.
REFERENCES
1. L. Yu, C.G. Koh, L.J. Lee, and. K.W. Koelling, Polym.
Eng. Sci., 42, 871 (2002).
2. C.-H. Wu and S.-H. Wu, J. Polym. Eng., 27, 107 (2007).

DOI 10.1002/pen

3. D.S. Kim, K.C. Lee, T.H. Kwon, and S.S. Lee, J. Micromech. Microeng., 12, 236 (2002).
4. A. Cramer, W. Michaeli, W. Friesenbichler, and I. uretek,
Zeitschrift Kunststofftechnik (WAK), 3, 1 (2007).
5. V.-W. Wang, Dynamic Simulation With Graphics for the
Injection Molding of Three-Dimensional Thin Parts, PhD
thesis, Cornell University, New York (1985).
6. M. Koponen, J. Enqvist, T. Nguyen-Chung, and G. Mennig,
Polym. Eng. Sci., 48, 711 (2008).
7. R. Blum, Verbesserte Temperaturkontrolle beim Kunststoffspritzgieen, PhD thesis, RWTH Aachen University, Aachen
(1993).
8. K. Ainoya and O. Amano, ANTEC, 726 (2001).
9. C.J. Yu, J.E. Sunderland, and C. Poli, Polym. Eng. Sci., 30,
1599 (1990).
10. D. Delaunay, P. Le Bot, R. Fulchiron, J.F. Luye, and G.
Regnier, Polym. Eng. Sci., 40, 1682 (2000).
11. R. Brunotte, Die thermodynamischen und verfahrenstechnischen Ablaufe der in-situ-Oberachenmodizierung beim
Spritzgieen, PhD thesis, Chemnitz University of Technology, Chemnitz (2006).
12. J. Greener and R. Wimberger-Friedl, Eds., Injection Molding for Microuidics Applications, Precision Injection
Molding Process, Materials, and Applications, Chapter 8,
169, Hanser, Munich (2006).
13. B.O Rhee, C.A Hieber, and K.K Wang, ANTEC, 496
(1994).
14. H.L. Zhang, N.S. Ong, and Y.C. Lam, Polym. Eng. Sci., 47,
2012 (2007).
15. D. Yao and B. Kim, J. Micromech. Microeng., 12, 604 (2002).
16. R.L. Golf, G. Poutot, D. Delaunay, R. Fulchiron, and E.
Koscher, Int. J. Heat Mass Tran., 48, 5417 (2005).
17. J.F. Dijksman and G.D.C Kuiken, Eds., IUTAM Symposium
on Numerical Simulation of Non-Isothermal Flow of Viscoelastic Liquids, Kluwer, Dordrecht, (1995).
18. D.C Venerus, J.D Schieber, V. Balasubramanian, K. Bush,
and S. Smoukov, Phys. Rev. Lett., 93, 0983011 (2004).
19. S.C. Dai and R.I. Tanner, Rheol. Acta, 45, 228 (2006).
20. T. Nguyen-Chung, G. Juttner, T. Pham, G. Mennig, and M.
Gehde, Zeitschrift Kunststofftechnik (WAK), 4, 1 (2008).
21. N.N., Moldow Plastics Insight, Online manual, Moldow
Corp., Framingham, MA (2008).
22. G. Juttner, Kunststoffe, 94, 53 (2004).
23. P. Kennedy, Flow Analysis of Injection Molds, Hanser,

POLYMER ENGINEERING AND SCIENCE-2010 173

Potrebbero piacerti anche