Вы находитесь на странице: 1из 8

Analysis of vacuum bag resin transfer molding process

M.K. Kang
a
, W.I. Lee
a
, H.T. Hahn
b,
*
a
Turbo and Power Machinery Research Center, Seoul National University, Seoul 151-742, South Korea
b
Department of Mechanical and Aerospace Engineering, University of California, Los Angeles, CA 90095-1597, USA
Received 10 April 2000; revised 27 November 2000; accepted 27 December 2000
Abstract
An analytical model is developed to analyze the resin ow through a deformable ber preform during vacuum bag resin transfer molding
(VBRTM) process. The force balance between the resin and the ber preform is used to account for the swelling of ber preform inside a
exible vacuum bag. Mold lling through multiple resin inlets is analyzed under different vacuum conditions. The formation of dry spots is
demonstrated in the presence of residual air. Molding of a three-dimensional ship hull with lateral and longitudinal stiffeners is simulated to
demonstrate the applicability of the model. q 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Vacuum; A. Resins; A. Fibers
1. Introduction
Resin transfer molding (RTM) is an effective method of
manufacturing composite parts of complicated shapes.
However, one major problem in applying the RTM process
to large parts is the tooling. It is difcult to build a closed
mold for large structures.
In vacuum bag resin transfer molding (VBRTM), an open
mold is used to hold the part that is sealed inside a vacuum
bag. Resin is introduced into the preform under vacuum
from inlet gates at ambient pressure, Fig. 1. The combina-
tion of an open mold and a vacuum bag reduces the tooling
cost and facilitates fabrication of large structures. The use of
vacuum reduces the void content in the composite and
improves the quality of the nished part as well [1].
However, since VBRTM is rather limited in the maximum
inlet pressure that can be used and in the maximum ber
compaction that can be achieved, its processing conditions
need to be controlled carefully.
The degree of vacuum in the vacuum bag determines the
amounts of air entrapped in the pores of the ber perform.
This residual air results in dry spots and micro voids. These
defects are in turn responsible for the degradation of
mechanical properties of the resulting part [28]. A spatial
variation of resin pressure inside the vacuum bag during
resin inltration may result in a non-uniform ber volume
fraction [9]. Since the resin pressure is higher near the inlet,
the ber preform will swell more there. Nevertheless, an
extended evacuation after complete mold lling can restore
most of the initial ber compaction by squeezing out excess
resin.
Since the inlet pressure is limited to ambient pressure, the
resin may not be able to ow fast enough to cover the entire
area especially if the structure is large. To speed up the mold
lling before the resin gelation, line gates may be preferable
to point gates. Needless to say, a proper design of injection
gates and evacuation ports is crucial for an efcient mold
lling. High permeability layers (HPL) can also facilitate
the resin ow by increasing the effective permeability [10].
The present paper describes the development and appli-
cation of simulation software for VBRTM. The deformation
of the ber preform during mold lling is accounted for by
using a unied force balance model. The compression effect
of the residual air is included to estimate the pressure build-
up during mold lling. Removal of excess resin to increase
the ber volume fraction is also analyzed.
2. Analysis
2.1. Governing equations
A unied model to describe the combined mass and
momentum balance between the resin and bers is given
by [11,12] (Appendix A)
2
1
V
f
2V
f
2t
2
1
V
f
2V
f
2x
i
v
f
i

2
2x
i
K
ij
m
r
2p
2x
j
_ _
; 1
Composites: Part A 32 (2001) 15531560
1359-835X/01/$ - see front matter q 2001 Elsevier Science Ltd. All rights reserved.
PII: S1359-835X(01)00012-4
www.elsevier.com/locate/compositesa
* Corresponding author. Tel.: 11-310-206-8157; fax: 11-310-825-2383.
E-mail address: hahn@seas.ucla.edu (H.T. Hahn).
where V
f
is the ber volume fraction, v
f
i
the velocity of
bers, p the resin pressure, m
r
the resin viscosity, and K
ij
is the permeability tensor of the anisotropic ber preform.
When high permeability layers are used over the ber
preform, an effective permeability tensor is used for K
ij
.
Subscripts i and j are the direction indices.
In VBRTM process, the resin ow is slow and the
preform deformation is negligible in the planar directions.
Therefore, the in-plane ber velocities are neglected. Since
a VBRTM part is usually thin compared to its planar dimen-
sions, the resin pressure and ber volume fraction are
assumed uniform through the thickness:
v
f
x
v
f
y

2p
2z

2V
f
2z
< 0: 2
The assumptions in Eq. (2) simplify Eq. (1) as
2
1
V
f
2V
f
2t

2
2x
i
K
ij
m
r
2p
2x
j
_ _
; 3
where indices i and j now denote x and y directions only, not
the thickness direction z.
Since the vacuum bag is exible, the ber deformation
stress and the pressure of resin or residual air balance ambi-
ent pressure, Fig. 2. The thickness and hence the ber
volume fraction of the preform are not uniform in the
plane because of the variation of pressure. Although other
equations are available, the CarmanKozeny equation is
used here to describe the permeability at an arbitrary ber
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1554
Nomenclature
A constant in the ber compaction model
d
f
diameter of ber
K
ij
permeability tensor
k
ij
Kozeny constants
n normal vector of control surface
n coordinate in the direction perpendicular to
the wall boundary
p pressure or resin
p
a
pressure of air
p
amb
ambient pressure
p
vac
vacuum pressure
s control surface
t time
v
f
i
ber velocity due to ber wash
V
f
ber volume fraction
V
f,0
ber volume fraction at zero compaction
pressure
x, y, z coordinates
Greek characters
l constant in the ber compaction model
m
a
viscosity of air
m
r
viscosity of resin
r
a
density of air
s ber compaction pressure
V control volume
Subscripts
1, 2 in-plane principal axes of ber preform
a air
amb ambient
CS control surface
CV control volume
f ber
i,j direction indices
r resin
vac vacuum
Superscripts
f ber
n current time step
n 21 previous time step
Fig. 1. Vacuum bag resin transfer molding process (VBRTM).
Fig. 2. Load sharing by ber perform and resin.
volume fraction [13]:
K
ij

d
2
f
16k
ij
1 2V
f

3
V
2
f
; 4
where d
f
is the ber diameter and k
ij
are constants.
The relation between the ber compaction pressure ( and
the ber volume fraction is proposed as
s A exp l
V
f
V
f;0
_ _
2expl
_ _
; 5
where V
f,0
is the ber volume fraction at zero compaction
pressure, and A and ( are model constants. Eq. (5) is a
modied version of the simple exponential model to satisfy
the zero-ber-compaction-stress condition at the inlet gate,
where V
f
V
f;0
: It is noted that the selection of the perme-
ability model and the ber compaction model is rather arbi-
trary since it does not affect the solution procedure. When
better models are available, Eqs. (4) and (5) can always be
replaced without affecting the basic solution procedure.
The residual gas inside the vacuum bag results from the
limited capacity of the vacuum pump used, defective seal-
ing, evaporation of volatile substances, etc. The maximum
vacuum would thus be obtained at the vapor pressure of the
volatile substances in the resin.
As illustrated in Fig. 2, the sum of ber compaction pres-
sure and the pressure of resin or gas should be equal to the
ambient pressure p
amb
p 1s p
amb
when the fiber preform is impregnated:
6a
p
a
1s p
amb
when the fiber preform is not impregnated:
6b
The equation relating the air ow to the changes in ber
volume fraction and pressure can be derived following the
same procedure as for Eq. (3) (Appendix A). The result is
given by
2
1
V
f
2V
f
2t
1
1 2V
f
p
a
2p
a
2t

K
ij
m
a
2
2
p
a
2x
i
2x
j
; 7
where m
a
is the air viscosity. The second term on the left side
is the result of the air being compressible. In the derivation,
the following observation has been used to simplify the nal
equation:
2p
a
2x
i
2p
a
2x
j
pp
a
2
2
p
a
2x
i
2x
j
: 8
The pressure distribution and the subsequent resin ow
inside the vacuum bag can now be calculated by using Eq.
(3) for the resin-impregnated area and Eq. (7) for the dry
area. The effects of temperature variation and chemical
reaction can be added to this model but they are not
accounted for in the present study.
As mentioned earlier, the current model should be modi-
ed for parts with a relatively large thickness compared to
the in-plane dimensions. However, many structures are
large in the planar dimensions and the present model
would offer a sufcient approximation.
2.2. Boundary conditions
The boundary conditions are chosen as follows:
1. Mold wall boundary: 2p=2n 0.
2. Case of continuous evacuation:
During mold lling
p p
amb
at inlet gates,
p
a
p
vac
at evacuation ports.
During removal of excess resin
2p=2n 0 at inlet gates (closed),
p p
vac
at evacuation ports.
Two different types of interface conditions are applied.
Where the ow front meets a dry spot, the resin pressure
is equal to the air pressure. If there is no dry spot, the ow
front pressure is set to the vacuum pressure.
3. Numerical solution scheme
The mold lling involves moving boundaries, demanding
a special treatment to move the calculation domain. In the
present study, a control volume nite element method
(CVFEM) is used to trace the ow front on a xed numer-
ical mesh, which consists of triangular elements and poly-
gonal control volumes [1418]. Integrating Eqs. (3) and (7)
within each control volume results in
For resin : 2
_
CV
1
V
f
2V
f
2t
dV
_
CS
K
m
r
7pn ds; 9a
For air :
_
CV
2
1
V
f
2V
f
2t
1
1 2V
f
p
a
2p
a
2t
_ _
dV

_
CS
K
m
a
7p
a
n ds; 9b
where CV is a control volume, CS a control surface, and n is
the normal vector of the control surface. Eqs. (9a) and (9b)
can now be transformed into matrix equations to be solved
numerically.
In the numerical formulation, the transient terms in Eqs.
(3) and (7) are discretized explicitly as follows:
2
1
V
f
2V
f
2t
2
1
V
n21
f
V
n
f
2V
n21
f
Dt
_ _
; 10a
1 2V
f
p
a
2p
a
2t

1 2V
n21
f
p
n21
a
p
n
a
2p
n21
a
Dt
_ _
; 10b
where superscript n denotes the current time step, and n 21
the previous time step. Dt stands for the time increment and
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1555
should be set sufciently small to guarantee a minimal
discretization error.
When solving Eq. (10b) on a xed grid system, difculty
is encountered because of the moving boundary. Since the
pressure of air in a dry spot is related to the volume of air, a
smooth numerical progression of the resinair interface is
essential for good prediction of the air pressure. An abrupt
change of the numerical ow front, which is inherent with
the conventional CVFEM, leads to a considerable error in
solving Eq. (10b). In the present study, an improved numer-
ical scheme called the oating imaginary nodes and
elements (FINE) method is employed for smooth progres-
sion of the ow front. In this method, an adaptive mesh-
rening technique is used to enhance the solution accuracy
at the ow front with only a marginal increase in the
computing time and memory [1921]. Thus, the mass
conservation of air is satised with less than 1% error.
4. Numerical examples and discussion
To illustrate the formation of dry spots, a square plate is
injected at its four corners under two different vacuum
conditions: continuous evacuation and initial evacuation.
In view of the symmetry only one quarter of the plate is
considered together with one inlet gate at a corner (Fig. 3).
A total of 800 uniform elements with 441 nodes are used to
cover the area.
4.1. Continuous evacuation
In the rst example, the vacuum pressure is set to the
vapor pressure of styrene (0.57 kPa) in the vinyl ester
resin. The material properties and the process conditions
are listed in Table 1. The simulation predicts the mold lling
time to be 1586 s. The inlet gate is sealed at the end of mold
lling, and the evacuation port starts draining excess resin at
the same vacuum pressure. The pressure drops rapidly at the
inlet gate and approaches the vacuum pressure (Fig. 4a).
Since the ambient pressure is balanced by the resin pressure
and the ber compaction pressure, the ber volume fraction
increases with decreasing resin pressure (Figs. 4b and c).
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1556
Fig. 3. Mold lling of a square plate injected at the corners: (a) pressure,
(b) preform thickness, (c) ber volume fraction, (d) mold ll fraction.
Fig. 4. Change of pressure, preformthickness and ber volume fraction with no dry spot: (a) pressure, (b) preformthickness, (c) ber volume fraction, (d) mold
ll fraction.
The time required to equalize the ber volume fraction
within 1% spatial variation is considerable (147 s). The
amount of drained resin from one quadrant of the rectangu-
lar array is 1.621 10
24
m
3
. The continuous evacuation at
the center eliminates the formation of a dry spot.
4.2. Initial evacuation
In the second example, the air is evacuated initially and
then the evacuation port is sealed. As expected, the pressure
of the entrapped air increases with the resin ow, Fig. 5a.
The initial vacuum pressure is set to the vapor pressure of
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1557
Table 1
Material properties and processing parameters used in the numerical simu-
lations
Effective values of d
2
f
=16k
11
and
d
2
f
=16k
22
m
2
(Eq. (4))
1.23 10
28
1.23 10
28
Fiber compaction model constants
A (Pa) (Eq. (5)) 5.138 10
25
l (Eq. (5)) 9.327
V
f,0
(Eq. (5)) 0.27
Resin viscosity, m
r
(Pa s) (Eq. (1)) 0.2
Vacuum pressure, p
vac
(Pa) 0.57 10
3
Ambient pressure, p
amb
(Pa) 1.01325 10
5
Fig. 5. Change of pressure, preform thickness and ber volume content with a dry spot at the center: (a) midship section of a ship hull (Sandown class
minehunter, UK), (b) mold lling pattern using point gates, dry spots not considered, (c) mold lling pattern using point gates, dry spots considered ll
time 1764 s, (d) mold lling pattern using injection channels, dry spots not considered.
Table 2
Material properties and processing parameters used for the three-dimensional ship hull
Woven berHPL stack
(for skin shell)
Same as in Table 1
Unidirectional berHPL
stack (for stiffener top)
Effective values of d
2
f
=16k
11
and d
2
f
=16k
22
m
2

(Eq. (4))
1.320 10
28
1.056 10
28
Fiber compaction
model constants
A (Pa) (Eq. (5)) 1.0 10
23
l (Eq. (5)) 5.0
V
f,0
(Eq. (5)) 0.2
Resin viscosity, m
r
(Pa s) (Eq. (1)) 0.2
Vacuum pressure, p
vac
(Pa) 0.57 10
3
Ambient pressure, p
amb
(Pa) 1.01325 10
5
styrene (0.57 kPa). With the resin ow and increasing air
pressure, the preform swells and the ber content decreases,
Fig. 5b and c. The fraction of mold area lled with resin is
shown as a function of time in Fig. 5d.
At t 7000 s, about 99% of the mold area is lled, and
the air pressure in the dry spot is about 99% of the ambient
pressure. The remaining 1% of the ambient pressure is borne
by the ber perform. Even such a low pressure can compress
the ber perform substantially since the perform is extre-
mely compressible initially, Eq. (5). This is the reason why
there is a large difference in ber volume fraction between
the inlet and the dry spot. However, it should be kept in
mind that such a large difference is an artifact of using the
one-dimensional equation for perform compressibility.
For a sufciently low vacuum pressure, as used in this
example, the nal size of the dry spot is negligible. Further-
more, the resulting dry spot may eventually break into
numerous microvoids because of the capillary effect.
4.3. Three-dimensional hull section
As a practical example, VBRTM of the midship
section of a battleship (Sandown class minehunter, UK
[22]) is analyzed. The overall dimensions are
13.06 m 5.45 m 7.62 m (WHD), employing 6180
non-uniform elements with 3402 nodes, Fig. 6a. High
permeability layers are also employed to increase the effec-
tive permeability. Two types of inlet gates are used: point
gate and channel gate. Injection channels are located at the
tops of the lateral and longitudinal hat stiffeners. The resin
can ow faster through the injection channels than through
the ber perform, and the mold lling time is reduced dras-
tically. The injection channels are assumed solid, and the
ow resistance in the channels remains constant. Since the
hat stiffeners contain unidirectional bers for better stiff-
ness, their permeability is anisotropic. The processing
conditions and material properties are summarized in
Table 2.
4.4. Continuous evacuation
In the case of point injection, 35 inlet gates and 50
evacuation ports are used. Fig. 6b shows the simulated
mold lling pattern when the continuous evacuation
prevents dry spots. The resin ows over the stiffeners and
into the skin. The time required for the mold lling is
3250 s. After the end of mold lling, an extra time of
378 s is needed to squeeze excess resin out, and the ber
volume fraction is improved to about 61% at the skin panels.
4.5. Initial evacuation
In this example, air is evacuated initially and the evacua-
tion ports are sealed. Because of the formation of dry spots
and the resultant pressure build-up of air, only 94% of the
mold is lled even after 4501 s (Fig. 6c). Filling the remain-
ing 6% requires an excessive amount of time, and eventually
will leave dry spots.
4.6. Injection channels
Using injection channels reduces the mold lling time to
as short as 1764 s in spite of the large dimensions (Fig. 6d).
When judiciously used, these injection channels can
substantially reduce the manufacturing time for large
structures.
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1558
Fig. 6. Mold lling patterns of a ship hull section using VBRTM with
different vacuum conditions.
5. Conclusions
While VBRTM is an attractive manufacturing process for
large structures, it requires a process optimization of higher
level than is necessary for conventional methods. In the
present study, a simulation model has been developed
which can account for the thickness change during resin
infusion. It can thus predict not only the mold lling time
but also the resulting part thickness. An additional feature is
its ability to predict the formation of dry spots containing
air. The simulation software developed in the present study
can be used to optimize the vacuum bag resin transfer mold-
ing process especially for large structures where trial and
error approaches can be quite expensive.
Acknowledgements
The present paper is partially based on work supported by
the Ofce of Naval Research Grant N00014-93-1-0650 with
Dr Yapa D.S. Rajapakse as the program director.
Appendix A. A unied model for the combined mass and
momentum balance between ber and resin
Assuming that the ber is perfectly impregnated by resin
leads to
V
f
1V
r
1; A1
where V
f
is the ber volume fraction and V
r
is the resin
volume fraction. The mass conservation of resin and ber
can then be written respectively as
2V
r
2t
1
2q
r
i
2x
i
0; A2a
2V
f
2t
1
2q
f
i
2x
i
0; A2b
where q
r
i
is the supercial resin velocity (ow rate per unit
cross-sectional area) and q
f
i
is the supercial ber velocity.
The actual velocities of resin and ber are thus given by
v
r
i

q
r
i
V
r
; A3a
v
f
i

q
f
i
V
f
: A3b
Combining (A1), (A2a) and (A2b) yields
2q
r
i
2x
i
1
2q
f
i
2x
i
0: A4
The relative velocity v
rel
i
between the resin and bers is
given by
v
rel
i
v
r
i
2v
f
i
: A5
In view of (A3a) and (A3b), (A5) can be rewritten as
q
rel
i
V
r

q
r
i
V
r
2
q
f
i
V
f
; A6
where q
rel
i
is the relative supercial velocity between the
resin and bers.According to Darcy's law, the relative
supercial velocity is related to the pressure gradient as
q
rel
i
2
K
ij
m
2p
2x
j
: A7
Combining (A6) and (A7) yields
q
r
i
2
V
r
V
f
q
f
i
2
K
ij
m
2p
2x
j
: A8
Taking the divergence of (A8), one has
2q
r
i
2x
i
2
2
2x
i
V
r
V
f
q
f
i
_ _
2
2
2x
i
K
ij
m
2p
2x
j
_ _
: A9
In view of (A1) and (A4), (A9) can be rewritten as
2
2q
f
i
2x
i
2
2
2x
i
1
V
f
21
_ _
q
f
i
_ _
2
2
2x
i
K
ij
m
2p
2x
j
_ _
: A10
Expanding (A10) leads to
2
2q
f
i
2x
i
2
1
V
f
2q
f
i
2x
i
1
1
V
2
f
2V
f
2x
i
q
f
i
1
2q
f
i
2x
i
2
2
2x
i
K
ij
m
2p
2x
j
_ _
:
A11
Using (A2b), one can rewrite (A11) as
2
1
V
f
2V
f
2t
2
1
V
2
f
2V
f
2x
i
q
f
i

2
2x
i
K
ij
m
2p
2x
j
_ _
: A12
In terms of the actual velocity, (A12) becomes
2
1
V
f
2V
f
2t
2
1
V
f
2V
f
2x
i
v
f
i

2
2x
i
K
ij
m
2p
2x
j
_ _
: A13
References
[1] LundstromTS, Gebart BR. Inuence from process parameters on void
formation in resin transfer molding. Polym Comp 1994;15(1):2533.
[2] Greszczuk LB. Effect of voids on strength properties of lamentary
composites. In: Proceedings of 22nd Annual Meeting of the Rein-
forced Plastics Division of the Society of the Plastics Industry,
1967;20.A-120.A-10.
[3] Hancox NL. The effect of aws and voids on the shear properties of
CFRP. J Mater Sci 1977;12:88492.
[4] Judd NCW, Wright WW. Voids and their effects on the mechanical
properties of composites an appraisal. SAMPE J 1978;14:1014.
[5] Yosida HT, Ogasa T, Hayashi R. Statistical approach to the relation-
ship between ilss and void content of CFRP. Comp Sci Technol
1986;25:318.
[6] Harper BD, Staab GH, Chen RS. A note on the effects of voids upon
the hygral and mechanical properties of AS4/3502 graphite/epoxy. J
Comp Mat 1987;21:2809.
[7] Feldgoise S, Foley MF, Martin D, Bohan J. The effect of microvoid
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1559
content on composite shear strength. In: 23rd International SAMPE
Technical Conference, 1991;259273.
[8] Bowles KJ, Frimpong S. Void effects on the interlaminar shear
strength of unidirectional graphite-ber-reinforced composites. J
Comp Mat 1992;26(10):1487509.
[9] Shenoi RA, Wellicome JF. Comp Mat Maritime Struct Vol. I: Fund.
Asp. 1993:10424.
[10] Moon Koo Kang, Thomas Hahn H, Resin transfer molding for naval
structures, US-Pacic Rim Workshop on Composite Materials for
Ships and Offshore Structures, 6-11 April, 1998.
[11] Dave R. A unied approach to modeling resin ow during composite
processing. J Comp Mat 1990;24:22411.
[12] Kempner, EA. Process simulation for manufacturing of thick compo-
sites, Ph.D. Thesis, University of California, LA, 1997.
[13] Dullien FAL. Porous media uid transport and pore structure. New
York: Academic Press, 1979.
[14] Bruschke MV, Advani SG. A nite element/control volume approach
to mold gilling in anisotropic porous media. Polym Comp
1990;11:398405.
[15] Fracchia CA, Castro J, Tucker CL. A nite element/control volume
simulation of resin transfer mold lling. In: Proceedings of the
American Society for Composites, 4th Technical Conference,
1989;157166.
[16] Advani SG. Flow and rheology in polymer composites manufactur-
ing. Comp Mat Ser 1994;10.
[17] Lin R, Lee LJ, Liou M. Non-isothermal mold lling and curing simu-
lation in thin cavities with preplaced ber mats. Int Polym Process
1991;6(4):35669.
[18] Lee LJ, Young WB, Lin RJ. Mold lling and cure modeling of RTM
and SRIM processes. Comp Struct 1994;27:10920.
[19] Kang MK, Lee WI, Kim TW, Kim BS, Jun E-J. Numerical
simulation of resin transfer molding process. Proc ICCM-10 1995;
3:25360.
[20] Kang, MK. A numerical and experimental study on mold lling and
void formation during resin transfer molding, Ph.D. Thesis, Seoul
National University, Korea, 1997.
[21] Kang MK. A ow front renement technique for the numerical simu-
lation of resin transfer molding process. Comp Sci Technol
1999;59:166374.
[22] Shenoi RA, Wellicome JF. Composite materials in maritime struc-
tures, volume II: practical considerations. Cambridge: Cambridge
University Press, 1993.
M.K. Kang et al. / Composites: Part A 32 (2001) 15531560 1560

Вам также может понравиться