Anda di halaman 1dari 11

AIJSTPME (2010) 3(1): 35-45

A Slit Die Design for Casting Plastics Sheets


Arunworradirok S.
Polymer Research Center and Department of Mechanical and Aerospace Engineering, Faculty of
Engineering, King Mongkuts University of Technology North Bangkok, Bangkok 10800, Thailand
Kolitawong C.*
Polymer Research Center and Department of Mechanical and Aerospace Engineering, Faculty of
Engineering, King Mongkuts University of Technology North Bangkok, Bangkok 10800, Thailand
Email address: ckw@kmutnb.ac.th
Abstract
This research objective is to investigate a simple way to design a plastics coat-hanger die that conveys the
uniform outlet velocity and pressure of a power-law fluid across the die width. A coat hanger die with a
circular manifold and parallel die land for a power-law fluid is investigated. Two charts of dimensionless
groups in the die manifold and land are plotted from an analytic calculation compared with those from a
numerical computation using commercial software. These charts help an engineer to design a plastics die
casting sheet and film. The comparison between the analytic and numerical reveals that the analytic charts
can only be used for thin film die design, while the other can be used for a thicker sheet. A design example
shows that the die shape built from our charts is in agreement with previous design at the same deformation
rate.
Keywords: Coat-hanger die, Plastic sheet Casting, Slit die, Die design
1 Introduction
Plastics sheet productions have three important
processing steps: molten plastics preparation, sheet
formation, and solidification. The molten plastics is
established in an extrusion machine managing the
production capacity, then the plastics sheet is formed
by a slit die, and finally solidified in air or water
before entering a rolling system controlling the final
product thickness. Thus, the slit die governs the
plastics sheet approximated thickness and the final
product quality. In general, the slit die is composed
of two main sections [1, 2]: manifold and die land.
The manifold delivers the polymer melt across the
die width and the die land forms the polymer melt to
be a plastics sheet and performs the molten polymer
flowing uniformly across the die lip. In some cases, a
choker bar and flex lip [3, 4] are attached at the die
land end to equalize the molten-plastics flow across
the die width. The slit dies have three main designs:
Tee, fish tail, and coat-hanger die (see Figure 1). The
Tee die is the first and the simplest design that does
not contribute the molten-plastics flow inside the die

across the die width. Thus, the polymer melts flow


very fast at the middle and slow near the die edges.
The fish tail and coat-hanger dies are the developed
designs helping the flow more uniform across the die
lip. However, the fish tail die provides a very lengthy
die land and thus a cumbersome die. The coat-hanger
die is the most versatile design. It can be designed to
fit a compact production site.
Figure 2 shows the coat-hanger die with its manifold
and land without the choker bar. The manifold is a
curved and round tube tapered from the middle to the
die edges. The manifold curvature controls the die
land length and eventually the total size of the die. In
some cases, the manifold cross section may be
designed in trapezoidal, rectangular, triangular or eye
drop shape [5]. Sometimes, the manifold and land are
designed such that the polymer melt flows in only
half of the cavity, while the other half is flat for
readily machining and cost saving [6]. The design of
coat-hanger die is not only useful for thin sheet and
film productions, but also for pipe manufacturing and

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

35

Arunworradirok S. and Kolitawong C.


blow molding [1, 2, 7]. In those processes, the
polymer melt from the metering section in the
extruder screw can be uniformly distributed to form
an annular shape by using the coat-hanger die
centerpiece.

Na, S. Y., and Lee, T. [16], apply the optimization


technique incorporated with a 3-D model finite
element method (FEM) for a power law liquid flow
in a coat hanger die. However, this technique
expenses lot of computer machine running time since
their optimal design reaches at the 4th iterations or
more.

(a)

Rin

y
H

Figure2: shows the coat-hanger die with its manifold


and land without a choker bar

(c)

Figure 1: shows the slit die designs: (a) Tee die,


(b) fish tail die, and (c) coat-hanger die
A coat-hanger die design for a power law fluid was
previously analyzed by many researchers [8, 9, 10,
11]; however, the manifold curvature is very lengthy
and design adjustment may be necessary due to inline space limitation and cost reduction [12].
To achieve the design adjustment, Smith, D. E. et. al.
[13, 14] apply a numerical optimization incorporating
with 2-D Hele-Shaw flow simulation in the design
with minimum pressure drop and reduced velocity
variation across the die exit. The optimal die shape,
however, is very complex and difficult to
manufacture. Later, they have developed their
method by restrict the manifold curvature and land
width, i. e. fix the in plane die shape, and optimize
only the manifold and land heights [3, 15]. Lately,

36

y(l )

x
l

(b)

yO

P1

RO

PO s

R(l )

Moreover, Sun, Y. and Gupta, M. [17, 18]


numerically study the elongational viscosity effects
of a low-density polyethylene (LDPE) melt flowing
in a coat-hanger die. The elongational viscosity is
found to slightly affect the velocity at the die exit. In
addition, Qu, J. and Zhang, X. [19] examine a coathanger die under axial vibration extrusion. They
found that the axial periodic pressure wave help
improve the mechanical properties, such as tensile,
impact strength and youngs modulus, in
polypropylene sheet since the pressure pulsation
increases the polymer crystallinity.
Michaeli, W. [1, 2] describes Wortberg, J. and
Kirchner, H.s coat-hanger design for a laminar,
isothermal, and incompressible molten polymer flow
in the die without end and edge effects and
viscoelastic behaviors. They use a representative
shear rate and viscosity on both the manifold
and die land. The representative shear rate is a
constant value at which the non-Newtonian shear rate
equals to the Newtonian one. Since the representative
shear rate correlates the representative viscosity, the
representative viscosities in the manifold R and in
the die land S are the nominal viscosities through
any cross section area of the manifold and die land,
respectively.
Figure 3 shows the representative shear rate concept
in a capillary. In the figure, the shear rate of a nonNewtonian fluid ( non Newtonian ) flowing in a capillary

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

A Slit Die Design for Casting Plastics Sheets

V = const

4V
( )
r =
4 S
R
4V
rS R e0
Newt . (r ) = 4 r
R
y
(r ) Non Newtonian

rS

dP rS
.
=
dy 2
dP r
(r ) =
.
dy 2

Wall Newt < Wall non Newt

rs = Representative radius in the capillary


= Representative shear rate

= Representative shear stress

Figure 3: shows an idea of the representative shear rate in a capillary [1, 2].
The shear rate of the non-Newtonian fluid flowing in a capillary radius R matches that of
the Newtonian one at the representative radius rs.
radius R matches that of a Newtonian one ( Newt ) at
the representative radius rs. Thus, the volumetric
flow rate in the manifold at any given length l is,
l
VR ( l ) = VO
L

(1)

where VO is a half of the total volumetric flow rate in


the coat hanger die; L, the half-die width; and l, a
local coordinate measured from the die edge (see
Figure 2). When the shear rate along the manifold is
steady, the representative shear rate of the polymer at
the middle of the manifold, (RO ) , equals that at any
manifold radius, (R ) . This gives,
l
R(l ) = RO
L

(2)

where RO is the maximum manifold radius located at


the middle of the die. Thus, the total pressure drop
across the die is the combination of the pressure drop
through the manifold and the die land which is,
P(l ) =
L

8VO R (s )s
12 S (l )VO
y (l )
ds +
4
LH3
R (s )L

(3)

where H is the die lip thickness, or in other words,


the approximate polymer sheet thickness. For a
uniform velocity across the die outlet, the pressure

drop in the die must constant along the die width.


That is,
(P )
=0
l

(4)

Then, Wortberg, J. and Kirchner, H. found that the


manifold curvature y(l) is,
l
y (l ) = y O
L

(5)

where yO, the maximum die land located at the die


center, is
yO =

R H 3 L2
S RO4

(6)

If the representative viscosities in the manifold R


and in the die land S are the same, then Eqs. (2) and
(6) give,

RO = 0.482 BH 2

yO = 1.474 B 2 H

1
3

1
3

(7)
(8)

where the total die-lip width,


B = 2 L

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

(9)

37

Arunworradirok S. and Kolitawong C.


Wortberg, J. and Kirchner, H. proposed a scheme
shown in Figure 4 to design a coat-hanger die which
can be used for the operating condition independence
on the left path. They examine the pressure drop
across the die by substitute l = L in Eq. (3) to get,
P =

12VO S yO
LH3

(10)

and the residential time,


t=

L H yO
VO

< tD

(11)

where tD is the degradation time of the polymer melt.


Moreover, Michaeli, W. also informs that Grmar
correlates the flow rate VO and the pressure drop
across the die P by two dimensionless sets; one for
the flow in the manifold and another one for that in
the die land,

Figure 5: illustrates Grmars dimensionless charts


to design a coat-hanger slit die for Prandtl-Eyrings
hyperbolic polymers [1].
4VO
1 P RO 1 P RO

3
ROC 3 A L 3 A L

6VO
1 H P 1 H P

=
H 2C L 2 yO A 2 yO A

(12)

(13)

where (u ) and (r ) are hyperbolic functions


defined such that,

(u ) =

(r ) =

1
1
8 1

cosh u sinh u + 2 (cosh u 1)


u 2 2
u
u

3
1

cosh r sinh r
r
r 2


= C sinh
A

(14)
(15)
(16)

A and C in the Eqs. (12) and (13) are constants of the


Prandtl-Eyrings hyperbolic constitute equation,

Figure 4: shows Wortberg and Kirchners flow chart


to design a coat-hanger die [Error! Bookmark not
defined.].

38

From Eq. (12)-(16), Grmar plots two dimensionless


charts for a coat-hanger die design shown in Figure
5. Since Grmar uses the Prandtl-Eyrings hyperbolic
constitutive equation, his slit die design is only useful
for the specific polymer. However, in general,
engineers design a slit die for the shear thinning
(power-law) plastics. Moreover, in many cases, the
total die length is limited because of space restriction
and cost saving, thus redesign is a must. Here, we
concentrate ourselves on the steady, laminar,
incompressible, power-law plastics flow in a coathanger slit die with a comfortable redesigned
scheme.

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

A Slit Die Design for Casting Plastics Sheets


Table 1: shows the shear rates of the Newtonian and
power law fluids axially flowing in a round pipe and
rectangular slit

Similarly, the representative viscosity in the die land


is,

S =

n
k 1
2VO
+ 2
2
3 n
LH

n 1

(21)

Thus, one can determine the maximum die land yO


for the power law fluid from Eq. (6). Moreover, from
Eq. (3), one can integrate the pressure drop in the die
to get,

2 Power law fluids


Since a desired slit die must have a uniform pressure
drop along the die width, a proper and simple coathanger die design can help a design engineer to save
time on the various redesigns to fit the factory needs.
Here, we model a coat hanger die for a steady,
laminar, incompressible, power-law polymer flow in
a slit die. Table 1 shows the shear rates of the
Newtonian and power law fluids flowing in a round
pipe and slit [10, 20, 21]. For a round pipe of radius
R in Figure 3, the representative radius rs at which
the shear rate of the non-Newtonian fluid
(non Newtonian ) equals that of the Newtonian one
( Newt ) is [22, 23],
1 1
n 1
RO + 3

4 n
n

=
rs

(17)

and the representative height, ys, in the slit is,


ys

1 1
n 1
3 n + 2

H
2

(18)

For a power law fluid, the viscosity is a function of


the rate of deformation,

= k n 1

(19)

where k, a consistency index, and n, a power law


index, are constants for a specific polymer and
temperature. Thus, substitute the power law shear
rate of a round pipe in Table 1 at the representative
radius rs [Eq. (17)] into Eq. (19) to get the
representative viscosity in the manifold,

R =

k 1
V
+ 3 O 3
4 n
RO
n

2
12V L13 2

3

L 3 l 2 3 + 12 S VO y l
P = O R4


O
3

LH L
RO


(22)
From the pressure drop of a power-law fluid flowing
in the die, Eq. (22), we propose dimensionless
relations [24, 25, 26] between the pressure drop in
the die and the flow rate in the manifold as,

P L2

1 2

R
2

VO RO
= R
,
L L

(23)

and similarly, in the die land,

P H 2

1 2

S
2

VO H
= S
,
L y

(24)

Then, from Equation (22)-(24), we can plot relations,


in the manifold, between the dimensionless pressure
drop across the die versus the dimensionless volume
flow rate for the dimensionless maximum manifold
radius, RO/L, from 0.01 to 0.05, and at the same time,
we plot those, in the die land, for the dimensionless
slit thickness, H/yO, from 0.01 to 0.2 as shown in
Figure 6. Here, we plot by using a typical HDPE at
453 K (180C) which has the consistency index, k,
6,190 N sn/m2, the power law index, n, 0.56 and the
density, , 950 kg/m3 [11]. Figure 7 shows a novel
coat-hanger die design schematic for the power law
fluid.

n 1

(20)

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

39

Arunworradirok S. and Kolitawong C.

(a)

Figure 7: shows a novel schematic to design a coathanger die for a power law fluid
3 Numerical comparison

(b)
Figure 6: (a) shows a plot [from Eq (22)] between
the dimensionless pressure drop versus the
dimensionless volume flow rate in the manifold when
the RO/L from 0.01 to 0.05 and (b) those in the die
land when the H/yO from 0.01 to 0.2

Numerical simulation is a cheap and reliable method


to validate the real flow in the slit die. Fluent, a
commercial finite-volume program [27], is capable
for simulating a non-Newtonian power law fluid
flowing in a 3-D model slit die [28]. We assume that
the flow is steady, laminar, incompressible and fully
developed. With the mass and momentum
conservation laws,

(25)
v dA = 0
and

40

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

A Slit Die Design for Casting Plastics Sheets

v v dA = p I dA + dA + F dV

(26)

where is the fluid density; v , the velocity vector;

dA , the cross sectional area normal to the flow; ,

the shear stress tensor; I , the identity tensor; V , the

dimensionless maximum manifold radius, RO L , and


dimensionless die-lip thickness, H yO , by
Percent Error =

1
N

PNumerical ,i PAnalytics ,i

i =1

PNumerical ,i

100%

(30)

fluid volume and F , the external forces. For


generalized Newtonian fluid, the shear stress tensor
is defined as

= ( )

(27)

where is the rate of deformation tensor and its


components are defined such that,

ij

u j ui
=
+

xi x j

1, 2,3
, i, j =

(28)
Figure 8: shows a computer simulation model with
its boundaries

where ui , u j are the velocity components in the i and


j directions, respectively. Also, xi , x j are the

distances in the i and j directions. The viscosity, , is


a scalar function of the shear rate which, in shear
flow, is the second invariance of the rate of
deformation tensor [9],

( )

1
:
2

(29)

As long as the polymer melt behaves like a shear


thinning material, the power law viscosity is
assumed, and defined in Eq. (19).
3-D extrudates of the die are constructed for only a
quarter of its entire shape to save the meshing and the
calculation time [29]. Figure 8 shows the model and
boundary conditions used in the numerical
simulation. Here, we use tetrahedral elements in the
manifold and hexahedral elements in the die land as
shown in Figure 9. Model details are in Table 2. With
adiabatic die walls, we vary ten values of the mass
flow inlet shown in Table 3 on each model to plot the
relation graph in Figure 10.
4 Discussions
From the dimensionless flow characteristics of a
power-law polymer flowing in the designed slit dies
by the analytics calculations in Figure 6 and the
numerical simulations in Figure 10, we compare the
pressure drops, P , in the dies at various

Figure 9: illustrates the meshing in the simulation


model: (a) Tetrahedral in the manifold and (b)
Hexahedral in the slit

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

41

Arunworradirok S. and Kolitawong C.


Table 2: shows the twenty-five die dimensions used
in the computer simulations; (a) varies the
dimensionless maximum manifold radius, RO/L, and
(b) varies the dimensionless die-lip thickness, H/yO.

only suitable for thin film die design with the H yO


recommended for 0.04 or less; however, the results in
Figure 10 is accurate for both thin film and sheet die
designs.
Table 3: illustrates the mass flow inlet used in the
computer simulations.
Number

Mass flow rate


(kg/s)

0.001389

0.002778

0.004167

0.005556

0.006944

0.008333

0.009722

0.011111

0.012500

10

0.013889

5 Design example

where PAnalytics ,i and PNumerical ,i are the pressure


drops in the slit dies from the analytical solution [Eq.
(10)] and the computer simulation, respectively.
Figure 11 shows the comparisons. In the figure, the
pressure drops from the analytics give about 13
percent overestimate the simulation independent to
the RO L . Moreover, the higher H yO , the higher

An engineer would like to design a coat-hanger die to


produce Polypropylene (PP) plastics sheets of 2-m
wide and 0.005-m thick with an approximated
volumetric flow rate at 38.58 10-6 m3/s (250 kg/hr.).
He would like to use an extruder in his factory that
has the maximum pressure, Pmax , at 75 MPa. The
Polypropylene operated at the melting temperature
190 C has the consistency index, k, 48,900 N sn/m2,
the power law index, n, 0.41 and the density, , 900
kg/m3 [11].
To design a slit die, we follow the chart in Figure 7.
Since the plastics sheet is 2-m wide (B = 2 m or L = 1
m), and 0.005-m thick (H = 0.005 m), thus from Eq.
(7), we calculate the maximum manifold radius, RO,
at the middle of the die,

3
R=
0.482 (2 m) (5 103 m) 2 =
17.76 103 m.
O
1

pressure drops from the analytic solution departed


from those of the simulation. This contributes to the
lubrication approximation during simplification of
the analytic solution. Thus, the plot in Figure 6 is

42

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

A Slit Die Design for Casting Plastics Sheets

Percent Error

15
10
5
0
0.0200

0.0225

0.0250

0.0275

0.0300

0.0325

0.0350

0.0375

0.0400

R O/L

(a)

Percent Error

80
60
40
20
0
0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1

0.15

0.2

H /y O

(b)
(a)

Figure 11: shows averaged percent differences


between (a) the pressure drop in the die from
computer simulation, PNumerical ,i , and that from Eq
(18), PAnalytics ,i , at various dimensionless maximum
manifold radius, RO/L, and (b) the pressure drop in
the die from computer simulation, PNumerical ,i , and
that from Eq (18), PAnalytics ,i , at various
dimensionless die-lip thickness, H/yO.

This yields the dimensionless maximum manifold


radius, RO/L, 17.76 10-3. We then calculate the
representative viscosity in the manifold R and in the
die land S from Eqs. (20) and (21) to get

R =15.41103 N s m 2 and S = 23.25 103 N s m 2


respectively. In addition, we can determine the
dimensionless volume flow rate,
VO
= 0.7174 10 6
R L

(b)
Figure 10: (a) shows a plot from the computer
simulations between the dimensionless pressure drop
versus the dimensionless volume flow rate in the
manifold when the RO/L from 0.01 to 0.05 and (b)
those in the die land when the H/yO from 0.01 to 0.2

Thus, from Figure 10(a), we read the dimensionless


pressure drop in the manifold,

P L2
1 2
R
2

160

Therefore, the pressure drop in the manifold is;

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

43

Arunworradirok S. and Kolitawong C.


160 (15.41103 N s m 2 ) 2
2 (900 kg m3 ) (1 m) 2

21.11 MPa

in which the corresponding dimensionless pressure


drop in the die land is,

P H 2
1 2
S
2

(900 kg

m 3 (21.11 MPa ) 5 10 3 m

1
23.25 10 3 N s m 2
2
= 1.76 10 3

(900 kg m ) (38.58 10 m s )
(23.25 10 N s m ) (1m)

and the dimensionless volume flow rate is;

VO
S L

= 1.49 10 6

From Figure 10(b), we read the H yO 0.02 to get


the slit die curvature,
yO

(5 10 3 m)
0.02

= 250 10 3 m

Since the operating pressure P =


21.11 MPa is less
than the machine maximum pressure ( Pmax ), the
designed slit die can operate in the designate
extruder. From Eq. (11), the resident time of the
molten plastics in the extruder is;
t

(1 m)(5 103 m)(250 103 m)


3
(38.58 106 m )
s

32.40 s

By half of a minute, the molten polymer in the


extruder is not degraded, yet. Thus, we can build a
coat-hanger slit die shown in Figure 2 by using the
above parameters.
6 Conclusion
To design a coat-hanger die for a power law
materials, the maximum manifold radius, RO, which
is located at the middle of the die influences the
manifold size, R(l), and its curvature, y(l), while the
extruder machine capacity such as the machine
maximum pressure and flow rate confine RO.
Moreover, RO is also limited by an operating space
specification, machining cost, and die weight or the
land length at the middle of the die, yO. In short,
increasing RO decreases the pressure drop in the die,
P, and the maximum die land, yO.
Dimensionless graph from the computer simulations
in Figure 10 can help an engineer to easily design a

44

coat-hanger slit die that has the dimensionless


maximum manifold radius, RO/L, ranking from 0.01
to 0.05 and the dimensionless die-lip thickness, H/yO,
from 0.01 to 0.2 for any power law polymer. With
the suggested flow chart in Figure 7, engineers can
easily design a coat-hanger die for material
independent and also repeatedly redesign the die to
suit the engineer needs.
Acknowledgments
The authors acknowledge King Mongkuts
University of Technology North Bangkok for
funding supports.
References
[1] Michaeli, W., 1992, Extrusion Dies for Plastics
and Rubber: Design and Engineering
Computations, 2nd ed., New York, Hanser
Publishers, 52-74, 129-152.
[2] Michaeli, W., 1984, Extrusion Dies: Design and
Engineering Computations, New York, Hanser
Publishers, 173-210.
[3] Wang, Q., 2007, Advanced analysis and design
of polymer sheet extrusion, Ph.D. dissertation,
The University of Missouri, Columbia.
[4] Klein, I, and Klein, R., 1973, SPE Journal; 29:
33-37.
[5] Wilson, G. M., and Garton, D. R., 1996, United
States Patent, No. 5,494,429.
[6] Appel, D. W., 1977, United States Patent, No.
4,043,739.
[7] Winter, H. H. and Fritz, H. G., 1986, Polymer
Engineering and Science, 26(8): 543-553.
[8] Chung, C. I. and Lohkamp, D. T., 1976.
Modern Plastics; 3: 52-55.
[9] Bird, R. B., Armstrong, R. C., and Hassager, O.,
1987, Dynamics of Polymeric Liquids: Volume
1, 2nd Ed., New York, John Wiley and Son, 170171, 245-246.
[10] Baird, D. G., and Collias, D. I., 1998, Polymer
processing: Principles and Design, New York,
John Wiley and Son, 19-20, 188-208.
[11] Tadmor, Z. and Gogos, C. G., 2006, Principles
of Polymer Processing, 2nd ed., New York, John
Wiley and Son, 706-711, 889-913.
[12] Matsubara, Y., 1981, United States Patent, No.
4,285,655.
[13] Smith, D. E., Tortorelli, D. A., and Tucker III,
C. L., 1998, Comput. Methods Appl. Mech.
Engr., 167: 283-302.

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

A Slit Die Design for Casting Plastics Sheets


[14] Smith, D. E., Tortorelli, D. A., and Tucker III,
C. L., 1998, Comput. Methods Appl. Mech.
Engr., 167: 303-323.
[15] Smith, D. E., and Wang, Q., 2005, Polymer
Engineering and Science, 45(7): 953-965.
[16] Na, S. Y. and Lee, T., 1998, The Korean
Journal of Rheology, 10(1): 38-43.
[17] Sun, Y. and Gupta, M., 2003, Int. Polym. Proc.
Journal, 18: 356-361.
[18] Sun, Y. and Gupta, M., 2005, SPE ANTEC
Tech., 81-85.
[19] Qu, J. and Zhang, X., 2008, Journal of Applied
Polymer Science, 107: 1006-1019.
[20] Agassant, J.-F., Arenas, P., Sergent, J.-Ph., and
Carreau, P.J., 1991, Polymer Processing
Principles and Modeling, New York, Hanser
Publishers, 1-40.
[21] Bird, R. B., Stewart, W. E., Lightfoot, E. N.,
1960, Transport Phenomena, New York, John
Wiley and Son, 34-70.
[22] Arunworradirok, S., 2004, Parameter Analysis
for Designing a Cothanger Die, Master Thesis,
Department of Mechanical Engineering, King

[23]

[24]
[25]

[26]

[27]

Mongkuts Institute of Technology North


Bangkok. (in Thai).
Kolitawong, C. and S. Arunworradirok, 2005,
Journal of King Mongkuts Institute of
Technology North Bangkok, 15(4): 39-46. (in
Thai).
Panton, R. L., 1984, Incompressible flow, New
York, John Wiley and Son, 181-227.
Potter, M. C. and Wiggert, D. C. with Hondzo,
M., and Shih, T. I.-P., 2002, Mechanics of
fluids, 3rd. ed., Brooks/Cole, Thomson Learning
Inc., 237-263.
Sabersky, R. H., Acosta, A. J., and Hauptmann,
E. G., 1989, Fluid Flow: A first course in fluid
mechanics, 3rd ed., New York, Macmillan Inc.,
147-180.
Ferziger, J. H. and Peric, M., 2002,
Computational Methods for Fluid Dynamics, 3
ed., New York, Springer, 71-82.
Fluent User Manual, Fluent Inc., 2001.
Kolitawong, C. and S. Arunworradirok, 2009,
KMUTT Research and Development, 32(1): 322. (in Thai).
rd

[28]
[29]

King Mongkuts University of Technology North Bangkok Press, Bangkok, Thailand

45

Anda mungkin juga menyukai