You are on page 1of 9

Composites: Part B 57 (2014) 4755

Contents lists available at ScienceDirect

Composites: Part B
journal homepage: www.elsevier.com/locate/compositesb

Process simulation for a large composite aeronautic beam by resin


transfer molding
S. Laurenzi a,, A. Grilli a, M. Pinna a, F. De Nicola b, G. Cattaneo c, M. Marchetti a
a

Department of Astronautic Electrical and Energy Engineering, Sapienza Universit di Roma, Italy
Advanced Materials and Technologies Laboratory, Centro Italiano Ricerche Aerospaziali S.c.p.A., Italy
c
Research & Innovation Technology, Alenia Aermacchi S.p.A, Italy
b

a r t i c l e

i n f o

Article history:
Received 21 November 2012
Received in revised form 9 September 2013
Accepted 16 September 2013
Available online 26 September 2013
Keywords:
E. Resin transfer molding (RTM)
A. Polymermatrix composites (PMCs)
Numerical analysis

a b s t r a c t
This paper presents the numerical process analysis and the experimental investigation for the manufacturing of a reinforced carbon-ber demonstrator of a large aeronautic beam by resin transfer molding
(RTM). The component is a primary structure characterized by several thick sections with abrupt changes
in shape that complicates the resin impregnation of the preform. Process simulations based on a nite
element method-modied control volume (FEM-CV) were conducted to investigate the resin ow front
patterns and nd the injection scheme that guarantees both a good impregnation of the preform and a
lling time compatible with the resin gel time. The beam component was successfully manufactured,
and a good agreement between the numerical analysis and the fabrication process was demonstrated.
2013 Elsevier Ltd. All rights reserved.

1. Introduction
Aerospace industries always investigate new technology solutions responding to the market pressure and the technology demands. In order to comply with the environmental directions on
CO2 emissions, the next global objective is to reduce 50% of consumed fuel by 2020 and a further 20% by 2025. These objectives
can be reached in several ways, the use of lightweight structures
being the most cost effective from the industrial point of view.
For these reasons, aerospace companies, which are traditionally
based on the use of metal alloys, have turned to the research and
development of composite polymeric materials. The main advantages of polymeric composites with respect to metals, such as resistance to corrosion and fatigue, and high performance/weight ratio,
are a set of qualities for winning the current and future applications. In this context, resin transfer molding (RTM) is a cost-competitive process to manufacture composite structures for
aeronautics [15]. Many parts manufactured using RTM in the
aerospace eld have been mostly related to non-critical structures,
whereas the development of large critical structures by RTM still
requires large research efforts. A good design for RTM leads to fabrication of three-dimensional near-net-shape parts, offering production of cost-effective structural parts in medium volume
quantities [2,6].
Also, so far RTM process have been used for the manufacture of
thin laminate structures and very little work on thick composites
Corresponding author. Tel.: +39 06 49 919 756; fax: +39 06 49 919 757.
E-mail address: susanna.laurenzi@uniroma1.it (S. Laurenzi).
1359-8368/$ - see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compositesb.2013.09.039

can be found in the literature. As a matter of fact, manufacturing


a thick-sectioned structure by RTM is challenging and several
important processing considerations have to be accounted for.
The highly exothermic nature of thermoset resins and the limited
temperature control make it difcult to avoid detrimental thermal
and cure gradients within the composite [7]. Moreover, the resin
transverse ow front, due to the preform permeability through
the section, becomes relevant. The resin ow patterns are difcult
to predict, making the gates and vents location analysis even more
difcult to perform [6]. In the case of thick-section composites,
therefore, the process features are a serious limitation for the manufacture of composite materials for critical structures. Process
modeling can accelerate the path from conception to prototype,
thus reducing industrial costs and time.
In this work, numerical and experimental studies were performed to manufacture an aeronautic beam demonstrator in composite material that is traditionally made by metal alloy. The
component under investigation is a beam of a thrust reverser,
which is a large primary structure with complex geometry
(Fig. 1). The composite beam was redesigned from the metallic
and thicknesses ranging from 3 mm to 33 mm. An inverse engineering approach was applied to determine the permeability values needed to perform the simulation process. Filling simulations
based on nite element-modied control volume method were
conducted in order to nd the injection scheme that guarantees
both a good impregnation of the preform and a lling time compatible with the resin gel time. The beam component was manufactured using a RTM mold designed according to the results of the
process simulations.

48

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

Fig. 1. Beam element of thrust reverser in nacelle (top); the beam is typically 1000 mm long and it has complex 3D shape with several thick sections as shown in CAD model
(bottom).

2. Materials and methods


2.1. Materials
Hexcel G0926 carbon ber reinforcements were used for the
permeability experiments and for the component fabrication. The
weave was a 5H Satin with areal weight of 360 g/m2. The Henkel
Epsilon 99900 binder was used for the preforming. The binder
was dispersed at 10 wt% loading with respect to the ber reinforcement, and the preform was maintained under vacuum at 120 C for
1 h.
The resins used for the experimental tests were the epoxy Hexcel RTM6 and the benzoxazine Henkel Epsilon II 99110, whereas
the beam demonstrator was manufactured using the benzoxazine
system. This resin was selected because of its low exothermic
behavior that guarantees a large processing window, thus avoiding
premature resin gelling that can occur as a consequence of the
large dimensions of the component. The two resin systems were
used in according to the process cycles recommended by the producer. Table 1 summarizes the processing conditions adopted for
the two resin systems during the experimental and numerical
investigations.
2.2. Manufacturing
The resin selected to manufacture the beam requires that the
entire apparatus (injection machine, dispensing resin and mold)

is pre-heated to 8090 C before starting the impregnation phase.


In our experiments, we used a commercial Hyperjet machine for
mono-component polymers, which injects the resin at a constant
pressure. The resin was loaded, degassed and heated inside the
RTM equipment. When the set temperatures of both the resin
and the mold were reached, the resin was pumped inside the mold
until it came out from the vent. After the ber preform was completely saturated with the resin, curing reactions were allowed to
continue past the gel point to form a cross-linked polymeric
structure.
Due to the complex geometry of the beam, several perform
pieces were assembled to achieve the nal part shape. Binder powder was applied to stabilize the layers. The lay-up of each part was
designed using the commercial FiberSim software, in order to
avoid ber angle deviation and waste of material during the preform preparation, and for the reproducibility of the process. The
plies were cut using an automatic machine and then placed in
the mold. Finally the preform is consolidated under vacuum.
Fig. 2 shows the process steps from the design to the perform fabrication. In particular, two consecutive ply creation steps are
shown in Fig. 2a and b, where the green lines represent the boundary layers. Fig. 2c shows the layer after the automated cutting and
the mold used to pre-shape the part, nally Fig. 2d shows a picture
of the consolidation phase.
The beam was designed with quasi-isotropic laminations considering constant ber volume fraction. The stacking sequence
adopted for the characteristic thicknesses of the preform is

Table 1
Processing conditions adopted for experimental tests and numerical analysis. The demonstrator was fabricated with the process parameter of the system Epsilon II 99110 as
reported in this table.
Resin

Density (g/cm3)

Pre-heating T (C)

Injection T (C)

Viscosity at Tinj (mPa s)

Gel time (min)

Cure cycle

HexFlow RTM 6
Epsilon II 99110

1.12
1.22

80
110

80
110

180
100

240
240

75 min at 160 C
90 min at 180 C

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

49

Fig. 2. Steps showing from concept to realization of a part of the preform; rst, the shape and paths of each layer is studied by simulating the laying and the overlapping (a
and b). When the ply-book is completed, the layers are cut by an automatic machine (c), nally the layers are positioned on the mold and the preform is consolidated under
vacuum.

reported in Table 2. The beam mold was designed in steel alloy


with an optimal distribution of electrical heating points to achieve
the uniform temperature required for the process. The heating system was controlled by a unit connected with thermocouples inside
the mold. During the injection phase, the temperature gradient

Table 2
Characteristics of stacking sequence adopted for the typical thicknesses of the
preform; these features were also used for the experimental part related to
permeability calculation.
Thickness
(mm)

Fiber volume
fraction

Number of
plies

Lamination sequence

3
10

0.55
0.55

8
26

20

0.55

52

30

0.550.53

76

[0/+45/45/90]S
[0/(03/+453/453/
90)]S
[0/(03/+453/453/
90)2/90]S
[0/(03/+453/453/
90)3/90]S

was also monitored by a thermal camera. In order to avoid race


tracking of the resin between the preform and the mold wall, the
ends of the preforming tools were sealed using silicon blocks. A
multipart mold design was adopted to allow for an easy demolding.
The RTM apparatus described above was also used to measure
the experimental lling time presented in Section 4.1. In this case,
the mold was a at metal tool with dimensions of
400 mm  400 mm, and the gap-height was adjusted in the range
between 3 mm and 30 mm by spacer frames. The resin was injected
throughout the mold using a circular inlet while the outlet vent was
positioned in opposite part. The inlet provided an internal linear
channel in order to produce a uniform preform impregnation. These
experimental parts had thicknesses specied in Table 2, using the ber reinforcement and the resins selected for the beam (Table 1) in
order to determine the lling time for the permeability analysis and
to investigate possible problems related to the use of the materials.
The thicknesses and the related lay-up sequence used for each plate
were identied as the most representative for the beam.

50

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

3. Flow analysis
The numerical analysis was performed using the commercial
code MoldFlow. The modeling of the resin ow allows investigation of the resin ow patterns during the impregnation process
and strategical designing of the gates and vents locations [8]. A
critical issue is the optimal lling of the composite part while
avoiding dry spots and premature gelling of the resin.
The RTM process is simulated using numerical methods based
on nite element-modied control volume (FEM-CV) analysis. This
method uses the porous media ow approach, solving the Darcy
equation coupled with the continuity and energy equations at
any instant during the lling process for given boundary conditions. The control volume consists of a fabric unit cell formed by
bers and porosity.
The continuity equation is derived from the mass balance of the
ow rate in the control volume:

@q
U  rq q r  U s
@t

where q is the actual density of the resin, that is the mass of resin
divided by the control volume containing both resin
and ber, and

may not be constant within the control volume; U is the resin velocity, and s are the sink effects due to the porous medium.
The Darcys law Eq. (2) describes the ow of a viscous uid
through an anisotropic, homogenous, porous medium:

1
0
K xx
ux
1
@ uy A 
@ K yx
l1  mf K
uz
zx

K xy
K yy
K zy

10 @P 1
K xz
@x
A
K yz A@ @P
@y
@P
K zz
@z

where l is the resin viscosity, vf is volume fraction of ber, and


1  vf is the porosity of the control volume that is occupied by the
resin. Eq. (2) is written in components along the three principal
 prole that is directly proporaxes, and gives the average velocity

tional to the permeability tensor K and the pressure gradient rP.
The boundary conditions are given by the part geometry, the pressure value (or ow rate) at gate and vent positions.
Usually, the numerical analysis is assumed to be isothermal in
order to reduce both the cost-time of the runs and the number of
physical parameters. In particular, this assumption greatly simplies the model, drastically reducing the computational time to
reach the convergence, which in our simulations took approximately 320340 s of CPU time. Further, in this case, the material
properties necessary for the simulation are limited to the resin viscosity and the permeability tensor.
The FEM model adopted for the analysis of the full scale beam
consisted of 35,042 dual-domain triangular elements, as shown
in Fig. 3. A sensitive mesh analysis performed on the optimal injection, showed that with this number of elements, the simulation
converges to constant values in terms of lling time. Increasing
the number of elements leads only to an increase in the computational time. For the set up where the mold is maintained at constant temperature for the entire impregnation phase, simulations
were carried out isothermally considering for the resin viscosity
value at the mold temperature as reported in the material datasheets (Table 1). Material properties, such as permeability value, ber volume fraction and part thickness, were assigned to each
element allowing for non-uniform properties and variable
thickness.
Regarding the process parameters, the value of the injection
pressure was considered constant at 3 bars for the entire numerical
analysis. This value was extrapolated from experiments conducted
on representative at laminates to measure the lling time for the
permeability determination. In particular, it was observed that an

Fig. 3. View of the FEM model adopted in the simulation process of the beam with
the indication of the regions for the injection scheme.

injection pressure above 3 bars produces ber washing phenomena and local reinforcement deformations.
The main objective of the simulations was to determine the
injection scheme that guaranteed a good impregnation of the beam
preform in agreement with the process window. In particular, the
total lling time of the mold had to be lower than the resin gel
time. This requirement is strongly inuenced by the injection strategy. For components with simple geometry, the choices for the
injection gate positions are limited and can be easily adjusted
using simple simulations. In our case, the beam geometry was
rather complex and the choice of the initial injection scheme was
fundamental to reduce the number of possible gate locations. For
this reason, we performed the ow analysis considering a potential
injection scheme established by taking into account some physical
and technological observations. First of all, the ow rate through a
porous medium drops during the impregnation phase, and in particular tends to decrease drastically in the proximity of abrupt geometrical changes such as T-shape edges and increased thicknesses.
In order to compensate for the pressure drop it would be necessary
to increase the injection pressure. On the other hand, this solution
can produce a local ber deformation and ber washing phenomena, which affect the quality of the nal product. Alternatively, the
use of schemes with multiple gates can be adopted, even though it
is not recommended for complex shapes. The running of several
ow fronts can create dry spots and introduce operational difculties, because each of them needs to be controlled and opened at the
right time. Based on these considerations, there was a clear preference for the injection channels scheme (Fig. 3). In particular, the
starting injection scheme was linear and positioned along the thick
part of the beam corresponding to the elongated edge. This would
allow for one single resin inlet port, thus simplifying the resin dispensing system. In order to compensate for the pressure around
the ribs on the bottom side, additional gates were considered on
the short edge. The vent points were positioned on opposite sides
and close to the torque box to push the resin through it. This initial
injection scheme was adjusted until satisfactory results in terms of
lling times and preform impregnation were reached. The optimal
injection scheme was then used to design the mold for the beam.
4. Results and discussion
4.1. Determination of permeability
In general, preforms show different permeability value in different directions due to the anisotropy of the ber architecture. As the
thickness of the composite part is typically negligible compared
with the in plane-dimensions, the permeability assumes a

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

two-dimensional form, where the in plane-directions ow resistances are considered predominant. On the other hand, in the case
of thick preforms, the permeability value in the out-of-plane direction cannot be neglected due to the ow front through the thickness. This value is difcult to determine experimentally. Several
authors relied on laborious mathematical models that often require a large amount of experimental data [2,917]. From an
industrial point of view, these approaches are not cost-effective
and a faster method to obtain input for the design of the injection
scheme is required. In order to overcome these issues, in our work
we introduced an effective in-plane permeability value, which also
takes into account the transverse ow through the thickness. This
approach started from the observation that the permeability K is
indirectly determined by experimental data. Usually, the in-plane
measurements consist of recording the ow front position at each
time, and then tting the data by the integral of the Darcy equation
between the position of the injection point at time t0 (x0, y0) and
the position of the ow front at time tf(xf, yf). The integral of the
Darcys law for the one-dimensional experimental approach is
given by the following expression:

xf

s
2K X rPt f

lu

where the viscosity l, the pressure gradient and porosity u are


known, the xf and tf are experimentally determined. The permeability value is calculated as the slope of the curve obtained from Eq.
(3):

Kx

lu x2f
2rP t f

The permeability values determined experimentally suffer generally from lack of reproducibility, and a large number of experiments are required to obtain an average value. We should stress
that typically the permeability is calculated from the linear part
of the curve assuming that the resin ow can be considered quasi-stationary. In reality, the curve can be not linear due to the
dual-scale porosity of the preform, and therefore the permeability
value is an approximation of several factors that relate to the different resin ows through the preform. An example of typical trend of
the squared ow front position (x2f ) as a function of the lling time
(tf) is presented in Fig. 4. In the graph, the dashed line is the linear
approximation of the curve used to determine the permeability. The
plotted data refer to the real case of a carbon ber-reinforced laminate for permeability experiments that were conducted using two
different injection pressures, as indicated in the picture. Starting
from these considerations, we considered a different approach
based on the simple measurement of the total lling time necessary
to impregnate a rectangular plate. The experimental set-up was
previously described in Section 2.2. After determining the preform
lling time through experiments, an iterative numerical analysis
was run by varying the permeability value until the same lling
time obtained experimentally was reached. The numerical analysis
was based on a FEM-CV model reproducing exactly the part geometry as the experiment, and with the material data (viscosity and ber volume fraction) as those adopted in the experiments. In this
approach, the permeability value is resolved along the total geometrical dimensions and the total lling time, and it includes the
interaction of ber reinforcement and resin that are used in reality.
In fact, the traditional experimental characterization of the permeability requires the use of a liquid (not resin) that interacts differently with the bers. In our case, the effects of the ow through
the thickness are included in a global permeability value. This
method is particularly suitable for quasi-isotropic laminates, where
the permeability values are almost the same. In our procedure the
in-plane permeability values Kxx and Kyy were assumed to be equal

51

Fig. 4. Example of typical curves of the square of ow front position xf as function of


the lling time tf obtained for two different injection pressure Pinj. The dashed line
represents the linear approximation used to calculate the permeability.

following the quasi-isotropic lay-up sequence preform. Following


the laminate characteristics summarized in Table 2, the permeability values that we determined with this global method were
6  1012 m2 for thickness of 30 mm, 3.6  1011 m2 for 20 mm,
4.73  1011 m2 for 10 mm, and 12.4  109 for 3 mm.
4.2. Simulation process
The ow analysis was used to simulate the mold lling process
during the impregnation phase. The resin ow front position for
each lling time, the pressure distribution at the end of the process, and the total mold lling time were the output of the simulations. From these simulations, potential macrovoids and problems
related to the impregnation of the preform can be extrapolated.
Three different injection schemes were selected based on the
considerations discussed earlier. The process conditions for the
simulations are the following:
 Case 1: linear injection channel along the entire edge of the
transverse beam.
 Case 2: linear injection channel along the length of the beam
except for the torque box area.
 Case 3: linear injection channel further reduced with respect to
simulation Case 2.
Results of the simulations show that varying the length of the
main linear channel strongly affects the resin ow front position
and the lling time of the impregnation step. Fig. 5 shows the case
where the injection scheme is positioned along the entire length of
the beam (simulation Case 1). Here, the ow front reaches the vent
position before nishing the impregnation of the torque box region, which remains dry because the resin ow circumvents this
zone. Upon reducing the length of the main channel as in simulation Case 2, the ow slows down in the torque box area and the
impregnation is guaranteed (Fig. 6). A further reduction of the
injection channel (simulation Case 3) causes a drastic pressure
drop around the torque box (Fig. 7). By comparing the ow front
position of simulation Cases 2 and 3 at the same time (Fig. 8), we
can observe that in simulation Case 3 the torque box region is still
not impregnated, unlike the situation in simulation Case 2. In simulation Case 3 to obtain a total impregnation of the preform, a time
of 3 h 40 min is necessary, whereas in simulation Case 2 the required time is about 2 h. Furthermore, as previously noted, this is
the result of a numerical analysis conducted under isothermal
assumptions. We have to stress that in the actual process the resin

52

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

Fig. 5. Simulation Case 1; numerical results of injection scheme with the injection channel along the thick edge and on the short one. In particular, the picture shows the ow
front position as a function of the lling time and the pressure gradient distribution inside the mold at the end of the lling. The total lling time is 1 h 54 min.

Fig. 6. Simulation Case 2; numerical results of the scheme with the injection channel along the thick edge having a reduced length respects to the Case 1. In particular, the
picture shows the ow front position as a function of the lling time and the pressure gradient distribution inside the mold at the end of the lling phase. The total lling time
is 2 h 6 min.

is exothermic and a signicant viscosity increase occurs during this


time. As reported in the material datasheet, the viscosity at 110 C
can be considered constant for about 3 h. After this time the viscosity increases, and at approximately 4 h, as in the simulation Case 3,
the viscosity becomes ve times the initial value. Therefore, the
actual time of the impregnation phase is larger than the numerical
one, with a high of probability to run into premature gelation of the

resin and the formation of dry regions due to the pressure drop.
Fig. 9 compares the ow patterns of simulation Cases 1 and 2 at
the same time step before reaching the vent. It is clear that in Case
1 the resin reaches the vent before impregnating the entire preform. The expected dry zones are indicated in Fig. 9.
Fig. 10 gives an enlarged view of the torque box area for the
cases considered here. In particular, the picture shows the ow

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

53

Fig. 7. Simulation Case 3; numerical results of the scheme with the injection channel along the thick edge having a reduced length respects to the Case 2. In particular, the
picture shows the ow front position as a function of the lling time and the pressure gradient distribution inside the mold at the end of the lling phase. The total lling time
is 3 h 40 min.

Fig. 8. Comparison of ow patterns for simulation Cases 2 and 3 at a given lling time.

front position at the lling time when the resin reaches the vent for
the injection scheme of simulation Case 1. In this case, a large dry
area is formed in proximity of the torque box by the time the resin
has reached the vent (Fig. 10a). This is due to fact that the injection
scheme is along the entire length of the beam, and the resin injection channel is near to the vent. Therefore, the resin reaches the
vent before nishing the impregnation of the entire torque box

area. In addition, the resin infuses fast through the


T-shaped edges with higher permeability circumventing the zone
with larger thickness. In simulation Case 2, the length of the injection channel goes up to the torque box region (Fig. 10b). In this situation, the resin has time to impregnate the preform and proceeds
uniformly towards the vent. Continuing to reduce the length of

54

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

Fig. 9. Comparison of ow patterns for Case 1 and Case 2 at a given lling time. In Case 1, the red arrows point to the expected dry zone caused by the resin reaching the vent
before the impregnation of the preform is completed.

Fig. 10. Comparison of the ow front position at the given lling time around the torque box areas for the three simulation cases under consideration. In simulation Case 1, a
dry area occurs in correspondence of the torque box when the resin has already reached the vent (a). In simulation Case 2, the resin ow front proceeds uniformly towards the
vent (b). In simulation Case 3, the torque box area shows a large dry zone when the resin is in proximity of the vent (c).

S. Laurenzi et al. / Composites: Part B 57 (2014) 4755

55

Fig. 11. View of the transverse beam manufactured by RTM.

injection channel, as in Case 3 (Fig. 10c), the torque box area shows
regions that are not completed lled.
Based on these analyses, the injection scheme of the simulation
Case 2 was adopted to design and fabricate the mold. This injection
scheme represents a good compromise between the process window and the material behavior. We found that the experimental
lling time to manufacture the beam was in agreement with the
numerical values. The injection time was about 2 h 20 min, very
close to the predicted value with an injection pressure of 3 bars.
The nal piece shows a complete impregnation of all the required
areas with no macroscopic defects on the surface (Fig. 11). It
should be underlined that the use of effective permeability values,
as we reported in this work, was an efcient way to reduce the
number of experimental investigations required for such a complex geometry, resulting in a good agreement between the experimental and numerical lling times.
5. Conclusions
In this work a lightweight structural beam was successfully
manufactured and prototyped in polymeric composite materials
as a replacement for aluminum based ones. The composite component has paintable and aesthetical high grade surfaces, and offers a
20% weight saving over the aluminum part. The injection scheme
was established following the numerical ow analyses while
adopting a new methodology to include the out-plane ow effects
typical of thick sections in effective in-plane permeability values.
This method is particularly useful in industrial settings where
experimental and trial error methods should be minimized. This
technological demonstrator is a proof of concept for resin transfer
molded structural composite parts to replace metal in primary
structures for aeronautic applications.
Acknowledgements
This project was funded by the Italian Ministry of Education,
University and Research (MIUR). The authors acknowledge DEMA
S.p.A. (Italy) for contribution to the design and construction of
the beam mold. Dr. M.G. Santonicola (MESA+ Institute for
Nanotechnology, University of Twente, the Netherlands) and

G. Attolini (Sapienza Universit di Roma, Italy) provided useful


comments on the manuscript.
References
[1] Laurenzi S, Marchetti M. Advanced composite materials by resin transfer
molding for aerospace applications. In: Ning H, editor. Composites and their
properties. Rijeka, Croazia: Intech; 2012. p. 127226.
[2] Kruckenberg T, Paton R. Resin transfer moulding for aerospace
structures. Netherlands and Norwell: Kluwer Academic Publishers; 1998.
[3] Laurenzi S, Di Nallo D, Marchetti M, Lalia Morra E, Anamateros E.
Manufacturing approach to realize a prototype of a helicopter transmission
component by composite materials. In: Proceedings of ERF31. Florence, Italy;
September, 2005. p. 168396.
[4] Laurenzi S, Marchetti M, Anamateros E. Liquid composite molding for
aeronautical application. In: Proceedings of 25 ICAS. Hamburg, Germany;
September, 2006. p. 256572.
[5] Soutis C. Carbon ber reinforced plastics in aircraft construction. Mater Sci Eng
A-Struct 2005;412(12):1716.
[6] Advani SG, Sozer EM. Process modeling in composites manufacturing. Marcel
Dekker Press; 2002.
[7] Ruiz E, Trochu F. Numerical analysis of cure temperature and internal stresses
in thin and thick RTM parts. Compos Part A-Appl S 2005;36(6):80626.
[8] Brouwer WD, van Herpt ECFC, Labordus M. Vacuum injection moulding for
large structural applications. Compos Part A-Appl S 2003;34(6):5518.
[9] Amico S, Lekakou C. An experimental study of the permeability and capillary
pressure
in
resin-transfer
moulding.
Compos
Sci
Technol
2001;61(13):194559.
[10] Barlow D, Howe C, Clayton G, Brouwer S. Preliminary study on cost
optimisation of aircraft composite structures applicable to liquid moulding
technologies. Compos Struct 2002;57(14):537.
[11] Luo Y, Verpoest I, Hoes K, Vanheule M, Sol H, Cardon A. Permeability
measurement of textile reinforcements with several test uids. Compos Part
A-Appl S 2001;32(10):1497504.
[12] Parseval YD, Pillai KM, Advani SG. A Simple model for the variation of
permeability due to partial saturation in dual scale porous media. Transport
Porous Med 1997;27(3):24364.
[13] Simacek P, Advani SG. A numerical model to predict ber tow saturation
during liquid composite molding. Compos Sci Technol 2003;63(12):
172536.
[14] Slade J, Pillai KM, Advani SG. Investigation of unsaturated ow in woven,
braided and stitched ber mats during mold-lling in resin transfer molding.
Polym Compos 2001;22(4):491505.
[15] Saouab A, Brard J, Lory P, Gardarein B, Bouquet G. Injection simulations of
thick composite parts manufactured by the RTM process. Compos Sci Technol
2001;61(3):44551.
[16] Turner DZ, Hjelmstad KD. Determining the 3D permeability of brous media
using the Newton method. Compos Part B-Eng 2005;36(8):60918.
[17] Samir J, Echaabi J, Hattabi M. Numerical algorithm and adaptive meshing for
simulation the effect of variation thickness in resin transfer molding process.
Compos Part B-Eng 2011;42(5):101528.

You might also like