Вы находитесь на странице: 1из 180

- i -

UNIVERSITY OF NOTTINGHAM
SCHOOL OF MECHANICAL, MATERIALS
AND MANUFACTURING ENGINEERING
EXPERIMENTAL CHARACTERISATION
OF THE VACUUM INFUSION PROCESS
By Agnes Ragondet
Thesis submitted to the University of Nottingham
for the degree of Doctor of Philosophy
Oct 2005
- ii -
Abstract
Vacuum infusion (VI) is a process widely used for the manufacture of large
composite structures. It involves a stiff mould half and a flexible membrane that
seals the cavity. Vacuum is used to consolidate the preform and atmospheric
pressure drives low viscosity resin into the mould cavity. The part is cured at room
temperature. In order to reduce infusion time, a high permeability medium is added
to the preform. This material modifies flow dynamics during preform infusion.
This thesis presents a characterisation and analysis of the vacuum infusion process.
The aim is to analyse VI in order to understand better flow phenomena. The first of
the five chapters introduces the work and gives a general presentation of the
vacuum infusion process and the study conducted. Chapter 2 presents an
experimental overview of VI and the parameters believed to be key for the process.
The next chapter focusses on the effect of lay-up, permeability distribution and
thickness on flow and preform impregnation. Chapter 4 extends the study of flow
phenomenon in VI to the characterisation of the distribution medium responsible for
the creation of through-thickness flow. Chapter 5 presents a design methodology
for the development of an integrated distribution medium and a modelling analysis
for flow through such a structure in order to study void formation in the preform.
Results from this study show that two competing flows occur in the preform during
preform infusion, an in-plane flow behind the flow front and a trans-plane flow at
the flow front. Due to a difference in permeability between the preform and the flow
enhancement medium a delay occurs between flow front on the upper and lower
side of the preform, called lead-lag. Study of lead-lag shows that the impregnation
of the preform is completed in three steps where both in-plane and trans-plane flow
are involved and lead the impregnation at different times during the infusion.
Analysis of the distribution medium shows that the latter can be modelled as an
open channel. The conformation of the vacuum bag to the material as well as the
nesting of distribution medium layers dictates its flow enhancement properties.
Finally the replacement of the actual set-up by a membrane with integrated
distribution medium is proposed. Flow modelling through an array of channels and
embedded cells shows that void formation in the preform can be predicted and is a
function of both channels and preform permeability as well as preform thickness.
- iii -
Ackowledgements
I would like to thank my supervisors Dr Robitaille, Professor Long and Professor
Rudd for making this thesis possible and for their precious help, advice and support.
A great thank you to Roger Smith, Paul Jones, Dave Smith, Geoff Tomlinson and
Brian Foster for their support in the workshops and laboratories.
I would like to acknowledge the Roadlite project partners Europrojects, Vosper
Thornycroft, Southfields coachworks and the funding partners EPSRC and DTR
(Department of Transport).
I would also like to thank my parents, brother, grandparents and goddaughters,
Marjory and Eva, for their great love and support.
Thank you to Magali, Jrome, Aurlie and Fabien, temporarily landlords, for
welcoming me to their place and for their enthusiasm and cheerfulness.
Finally I would like to thank Jrmy for his love, support, patience and
understanding.
- iv -
Contents
Abstract i
Acknowledgements ii
Nomenclature vi
Chapter 1
Introduction
1.1. Description of the vacuum infusion process ....................................... 1
1.2. Flow characteristics in VI ............................................................... 3
1.3. Background ................................................................................ 7
1.4. Theme of the study ..................................................................... 11
1.5. Objectives of the study ................................................................ 12
Chapter 2
General analysis of the process and structural parameters
2.1. Introduction............................................................................... 17
2.2. Background ............................................................................... 17
2.3. Experimental analysis: materials.................................................... 22
2.3.1. Reinforcement............................................................................ 22
2.3.2. Fluid......................................................................................... 22
2.3.3. Consumables ............................................................................. 23
2.4. Assumptions .............................................................................. 24
2.4.1. Bagging strategy......................................................................... 24
2.4.2. Fabric style................................................................................ 25
2.4.3. End of infusion ........................................................................... 25
2.5. Experimental analysis: results ....................................................... 26
2.5.1. Bagging strategy......................................................................... 26
2.5.2. Effect of fabric style..................................................................... 31
2.5.3. End of infusion ........................................................................... 38
2.6. Conclusions ............................................................................... 42
Chapter 3
Analysis of the lead-lag in vacuum infusion
3.1. Introduction............................................................................... 46
- v -
3.2. Background ............................................................................... 47
3.3. Assumptions .............................................................................. 51
3.3.1. Flow front recordings................................................................... 51
3.3.2. Creation of lead-lag..................................................................... 54
3.3.3. Control of lead-lag ...................................................................... 54
3.4. Experimental analysis: materials and method ................................... 56
3.4.1. Materials................................................................................... 56
3.4.2. Method ..................................................................................... 56
3.5. Experimental analysis: results ....................................................... 57
3.5.1. Creation of lead-lag in VI.............................................................. 57
3.6. Conclusions ............................................................................... 78
Chapter 4
Experimental analysis and modelling of the distribution medium
4.1. Introduction............................................................................... 82
4.2. Background ............................................................................... 82
4.3. Assumptions .............................................................................. 85
4.4. Permeability measurement of the distribution medium........................ 85
4.4.1. Materials................................................................................... 85
4.4.2. Experimental set-up .................................................................... 85
4.4.3. Results ..................................................................................... 88
4.5. 2D CFD analysis.......................................................................... 89
4.5.1. Geometry.................................................................................. 89
4.5.2. Results ..................................................................................... 91
4.6. 3D CFD analysis.......................................................................... 92
4.6.1. Geometry.................................................................................. 92
4.6.2. Results ..................................................................................... 93
4.6.3. Modelling the nesting effect........................................................... 95
4.7. Experimental analysis of the distribution medium .............................. 98
4.7.1. Materials................................................................................... 98
4.7.2. Experimental set-up .................................................................... 99
4.7.3. Results ................................................................................... 100
4.7.4. Comments............................................................................... 103
4.8. Conclusions ............................................................................. 104
- vi -
Chapter 5
Novel vacuum membrane design methodology
5.1. Introduction............................................................................. 106
5.2. Hypothesis .............................................................................. 106
5.3. Background ............................................................................. 107
5.4. Preliminary study...................................................................... 109
5.4.1. Vacuum membrane geometry...................................................... 109
5.4.2. Modelling ................................................................................ 110
5.5. Proposed plan for future work...................................................... 125
5.5.1. Modelling ................................................................................ 125
5.5.2. Experimental analysis ................................................................ 126
5.5.3. Extension to general applications ................................................. 129
5.6. Conclusions ............................................................................. 130
Chapter 6
Conclusions and future work
6.1. Conclusions ............................................................................. 133
6.2. Future work............................................................................. 135
Appendices
Appendix 1........................................................................................... 138
Permeability measurements..................................................................... 138
Appendix 2........................................................................................... 150
Materials characteristics.......................................................................... 150
Appendix 3........................................................................................... 152
Viscosity of fluids................................................................................... 152
Appendix 4........................................................................................... 154
Thickness distribution in laminates at the end of infusion............................... 154
Appendix 5........................................................................................... 159
Effect of outlet pressure on lead-lag .......................................................... 159
Appendix 6........................................................................................... 162
Effect of varying preform thickness on lead-lag............................................ 162
Appendix 7........................................................................................... 164
Cost analysis of re-usable rubber membrane............................................... 164
Appendix 8........................................................................................... 167
References in alphabetical order by author.................................................. 167
- vii -
Nomenclature
Upper case symbols
A Fibre volume fraction at a compaction pressure of 1 Pa -
H Total thickness of n fabric layers m
K Permeability m
2
P Pressure Pa
Q Volumetric flow rate m
3
.s
-1
S Surface m
2
V Volume m
3
U Voltage Volts
Lower case symbols
v Velocity m.s
-1
u Superficial velocity m.s
-1
t Time sec
n Number of layers -
x Flow front distance m
h Thickness m
g Acceleration (9.81) m.s
-2
r Radius m
z Height m
Subscripts
a Average -
de Density -
f Fibre -
ff Flow front -
g Gravity -
m Mould -
- viii -
w Wall -
ll Lead-lag -
Superscripts
B Compaction stiffening index -
l Number of layers -
Greek symbols
| Aspect ratio of fluid flow front -
c Porosity -
p Density -
u Viscosity Pas
Abbreviations
VI Vacuum Infusion
CFRM Continuous Filament Random Mat
CFD Computationnal Fluid Dynamics
PW Plain Weave
BOF Bleed Out Fabric
DM Distribution Medium
RTM Resin Transfer Moulding
VARTM Vacuum Assisted Resin Transfer Moulding
VBRTM Vacuum Bag Resin Transfer Moulding
VARI Vacuum Assisted Resin Injection
RIFT Resin Infusion under a Flexible Tooling
HDPE High Density Polyethylene
PVC Polyvinyl Chloride
RPM Round Per Minute
CMM Coordinate Measuring Machine
- 1 -
Chapter 1
Introduction
1.1. Description of the vacuum infusion process
A composite material consists of a fibre reinforcement and a matrix. The
reinforcement, of high strength and stiffness, is impregnated by the matrix, which
keeps the fibres in their geometrical position and transmits stress through the
component, while guaranteeing chemical resistance. Composite materials are used
in a wide range of applications from automotive, aeronautic and sport sectors to
construction and architecture. Different manufacturing processes exist depending on
the type of application, production rate and size of the components.
Vacuum infusion (VI) is a widely used moulding process for the manufacture of
large composite structures. Its popularity is partly due to the low cost of the tooling
and the environmental safety (the process eliminates more than 90 % of the volatile
organic compound emitted by unsaturated polyester resins [1]). In addition, low
operator involvement increases the repeatability of the process compared to open
mould techniques such as hand lay-up or spray-up and the components are of
relatively high fibre content, up to 60 % by volume.
VI is increasingly popular in the transportation, marine and wind power generation
industries. Thick, single skin laminates and sandwich structures are produced using
this method [2, 3, 4]. Components are traditionally made of glass fibre and
polyester resin, but now VI is also used for the manufacture of carbon fibre - epoxy
components dedicated to aeronautic and aerospace sectors [5]. For example 70 m
long wind turbine blades are infused by VI. Their sandwich structure is made of a
wooden core, which is enclosed in a glass fibre and epoxy resin laminate. This
technique is also used to produce 22 m long train roofs made of balsa core and
glass fibre - vinyl ester resin sandwich laminates. In another transport sector, boat
hulls, typically 16.40 m long by 4.50 m wide and 2.50 m high, are produced by
infusing sandwich structures made of balsa core and glass fibre with polyester resin
using the VI process.
- 2 -
VI is part of a family of moulding techniques called liquid composite moulding. It
involves the infusion of a low viscosity resin into a dry fibrous preform placed on a
stiff mould and covered by a flexible membrane. The process relies on a pressure
gradient to drive the resin into the mould and impregnate the preform. The inlet of
the mould is at atmospheric pressure and the outlet is under vacuum. A typical set-
up for VI is illustrated in Figure 1.1. Figure 1.2. shows illustrations of possible
reinforcements and typical consumables used in the process.
Fig. 1.1. VI experimental set-up. Details of the mould lay-up.
The main advantages of VI, compared to open moulding process such as hand lay-
up or spray-up, are higher fibre content and an improvement of the laminate quality
due to a better impregnation of the preform. Operator involvement is less critical,
leading to higher quality and consistency. Also existing hand lay-up - spray-up
moulds can be modified for use in VI. There is no limitation in part dimensions and
it is possible to produce components from 1 mm thick to more than 100 mm thick.
Fig. 1.2. Typical fabrics and consumables used in VI.
Furthermore, the process reduces styrene emissions and capital costs (no injection
Vacuum bag
Distribution medium
Sealant tape
Mould
Resin channel
Outlet
P
2
Inlet
P
1
Reinforcement
Bleed out fabric
Reinforcement [6] Bleed out fabric Distribution medium Vacuum bag
- 3 -
machine is required), compared to other moulding techniques such as RTM. It
reduces cycle times when compared to hand lay-up, and permits accurate
placement of the reinforcement layers. Finally, all types of cores and inserts can be
inserted and moulded in-situ.
However, the process is sensitive to leakage in the flexible membrane and a good
surface finish is only available on one side of the part. Complex geometries such as
sharp edges and thickness variations can disturb the flow [7]. When core inserts
are included, they must be made of closed cells so as not to absorb resin which
would increase the weight of the component. Finally disposable materials (tape,
pipes, valves, distribution medium, bleed out fabric) produce waste and mould set-
up requires significant man-hours.
Vacuum infusion is known under different acronyms, all describing methods based
on the impregnation of a dry reinforcement by liquid thermoset resin driven under
vacuum. VARTM (Vacuum Assisted Resin Transfer Moulding, [3, 8]), SCRIMP
(Seemann Composites Resin Infusion Moulding Process, [9]), VBRTM (Vacuum Bag
Resin Transfer Moulding, [10, 11]), VARI (Vacuum Assisted Resin Injection process,
[12]) and RIFT (Resin Injection under Flexible Tooling, [13]) are the most popular
terms to describe vacuum infusion processes. All involve basically the same
technology, with some patented to cover different elements of the process, and
others generic names. SCRIMP focuses on the use of a flow enhancement medium
to reduce infusion time for the moulding of skin and sandwichs structures. The RIFT
method involves the use of a semi-flexible membrane, typically a glass - polyester
composite shell, instead of a smooth vacuum bag as used in VBRTM. The latter is
used as a generic name to describe the vacuum infusion process. VARI was
developed by the Institute of Structures and Design of the German Aerospace
Center [12]. The process is not different from a general vacuum infusion process
and the set-up includes a highly permeable layer on one side of the preform as for
SCRIMP. Finally VARTM is described as a variant of the RTM process [3] but it is
also used as a generic name to describe the vacuum infusion process [8].
1.2. Flow characteristics in VI
In order to better understand the impregnation dynamics in VI lets compare it to a
more simple process such as RTM. The latter involves a preform placed in a mould
- 4 -
cavity of constant thickness as shown in Figure 1.3. The infusion is conducted under
imposed pressure or flow rate. Since the cavity thickness is constant, the
permeability of the preform remains constant during the infusion. In this example
the impregnation dynamics are equivalent to a one-dimensional rectilinear in-plane
flow. The latter is assumed to correspond to the flow of an incompressible fluid
through a porous media, as described by Darcys law:
dx
dP K
v

=
u c
1.1
Where v is the insterstital velocity, K the preform permeability, the porosity, the
fluid viscosity and dP/dx the pressure gradient over the flow distance.
An integration form of Darcys law shows time as a function of flow front distance
squared as described in Equation 1.2.
P
x
K
t
A

=
2
2
c u
1.2
Fig. 1.3. Flow front pattern in RTM
In VI the upper face of the mould is made of a soft membrane. A pressure
difference is applied between the inlet of the mould, connected to a resin container
under atmospheric pressure, and the outlet of the mould, connected to a pump
under vacuum. The preform is ready to be infused when the air has been extracted
Inlet
P
1 P
2
Outlet
Lower mould
Reinforcement
Flow front
Dry area
Wet area
Upper mould
- 5 -
from the cavity, the air tightness of the mould has been checked and the pump has
been adjusted to the required level of vacuum. Then the inlet is opened and the
resin impregnates the preform. During the infusion, the compaction of the preform
evolves locally with the pressure gradient and the latter evolves with flow front
position as shown in Figure 1.4. The impregnated part of the preform is submitted
to a non-uniform pressure distribution with atmospheric pressure at the inlet and
vacuum at the flow front. The compaction and the permeability of the wet area
varies with position and flow front progression. The thickness of the impregnated
part is not uniform; the preform is typically thicker at the inlet where the pressure
level is higher. The dry part of the preform is under uniform pressure equivalent to
the vacuum set at the outlet. However, Darcys law remains valid for flow in VI, if
combined with the continuity equation (see Equation 1.1, 1.3) and if thickness and
porosity distribution are known. Flow in the cavity is in-plane only and rectilinear (at
this stage of VI description it is assumed that the lay-up has homogeneous
permeability i.e no layers of different permeability).
0 = Vv 1.3
Fig. 1.4. Evolution of mould cavity pressure with flow front progression during VI.
Distance from inlet
Pressure
P
a
= atmospheric
Flow front position 1
Flow front position 2
X
1
t
1
X
2
t
2
X
0
t
0
Vacuum bag
Preform
Outlet
P
2
Inlet
P
1
Wet Dry
Mould
- 6 -
In VI time is a critical factor as lay-up, infusion and part release are generally slow
processes especially when producing large structures. A solution to reduce overall
manufacturing time is to decrease infusion time by adding a distribution medium to
the preform. The latter, usually a polymer mesh of polyamide or polyethylene, can
also be a layer of reinforcement of high permeability made of the same fibre as the
preform [9].
The distribution medium is usually placed on the upper face of the preform, as it
has to be released after infusion (not if the distribution medium is a layer of
reinforcement). When the distribution medium is a polymer mesh, a layer of bleed
out fabric is placed between the preform and the distribution medium. This fabric, a
thin layer usually of woven polyamide and with low surface density (100 g.m
-2
),
assists the removal of the distribution medium. Due to its high permeability the
distribution medium increases resin flow at the upper surface of the preform, thus
reducing impregnation time, up to 86 % [14]. However, its higher permeability
distorts the shape of the flow front [9, 14]. The resin flows through the distribution
medium first and then impregnates the reinforcement. Therefore a delay occurs
between the upper and lower faces of the preform where the side in contact with
the distribution medium is impregnated ahead of the other side [14-17].
Fig. 1.5. Creation of lead-lag in VI.
Vacuum bag
Distribution medium
Outlet Inlet
P
1 P
2
Lower mould
Flow front
Reinforcement
Lead-lag
- 7 -
This delay is known as lead-lag (see Figure 1.5). It is important to notice that the
use of a distribution medium in RTM would also result in a lead-lag. Table 1.1 shows
a summary of the different types of flow encountered in RTM and VI.
Table 1.1. Flow characteristics encountered in VI and RTM.
VI RTM
Upper mould Soft, deformable Stiff, non-deformable
Variations in preform
permeability during infusion
-
No distribution
medium
- Rectilinear in-plane
- No lead-lag
- Rectilinear in-plane
- No lead-lag
F
l
o
w
Distribution medium
- In-plane and trans-plane
- Lead-lag
- In-plane and trans-plane
- Lead-lag
1.3. Background
The VI process was first developed by Marco in 1950 as an alternative to RTM for
the manufacture of boat hulls. The aim was to reduce voidage and tooling costs
[13]. The set-up consisted of a solid male mould on which dry reinforcement was
laid, covered by a semi-flexible female tool used for consolidation and air tightness.
The concept was not popular compared to hand-lay up and spray-up methods which
were more reliable techniques to produce large structures. The publication of health
and safety regulations on styrene emissions at work in 1974 encouraged industries
to develop closed mould processes [13]. The semi-flexible female tool, first used by
Marco, was progressively replaced by a silicone membrane. Resin was poured into
the dry reinforcement before closing the mould. Vacuum was used to consolidate
the preform and resin, under atmospheric pressure, was driven into the tool. Due to
lower operator dependency, components were of higher quality than those obtained
using the hand lay-up or spray-up methods and production rates increased.
However, flow control was difficult due to the high viscosity of the resin system.
From 1980 the use of vacuum to drive the resin into the mould became popular,
keeping the use of a silicone membrane as the female tool. This improvement
resulted in components of higher fibre volume fraction due to the compaction of the
fabric under vacuum. In addition low viscosity resin systems were developed [13] in
order to enhance flow and reduce infusion time.
In order to reduce infusion times further, especially when impregnating low
permeability preforms, flow enhancement media made of grooves and channels
- 8 -
were continuously developed. Twenty four patents were issued between 1959 and
1995 [1-4]. One notable flow enhancement patented technique is the SCRIMP
method developed by Seemann for the moulding of single skin and sandwich
structures [9, 13, 14, 18]. The method is very similar to other VI techniques as the
dry reinforcement is placed on a stiff mould half and the cavity covered with a
polyamide bag while the resin is driven inside the mould under vacuum. The
distribution medium consists of a criss-cross pattern made of plastic monofilaments
which create channels for resin transport. It can also be a knitted or woven fabric,
the important factor being the creation of a continuous network of channels over the
area to infuse [9]. The US patent specification involves a distribution medium on
one side of the preform only whereas a later European patent application uses a
distribution medium on both sides. The latter case is not so advantageous as the
high permeable layer located on the stiff mould side cannot be released from the
component after cure. Therefore a good surface finish is not achievable and the
distribution medium does not contribute to improve the mechanical properties of the
part but only increases its weight due to resin absorption.
As mentioned previously the distribution medium modifies the impregnation
dynamics and the fluid flows mainly through the thickness of the preform.
Depending on the type of fabric infused, this trans-plane flow evolves very
differently [19]. Isotropic fabrics such as continuous filament random mat are prone
to uniform flow progression through the thickness of the preform. Oriented fabrics,
such as unidirectional, result in micro-flows which progress between the fibre
bundles. The latter results in a non-homogeneous flow front on the side of the
preform opposite to that of the distribution medium [19].
The resulting distance between the flow front observed on the bag and that
observed on the mould is called lead-lag, as shown in Figure 1.5. The length of the
lead-lag depends mainly on the permeability of both the distribution medium and
the reinforcement, as well as the thickness of the reinforcement. Flow visualisation
experiments in VI made by Sun [14] show that the infusion of a 3-layer glass fibre
mat of permeability 50 times lower than that of the distribution medium results in a
60 mm long lead-lag. The latter is nearly constant throughout the impregnation
except at the beginning where it decreases [20, 17]. One problem encountered with
the presence of a lead-lag is that the upper side of the preform, in contact with the
distribution medium, is fully filled before the rest of the part. If the difference in
- 9 -
permeability between the distribution medium and the reinforcement is too large,
lead-lag can reach large distances, typically from 100 mm to 200 mm. When the
upper side of the preform is fully filled (distribution medium side), this distance
remains dry on the lower side. Resin flows preferentially in the distribution medium,
which is fully impregnated and has a higher permeability. Therefore it takes much
longer for the remaining dry area to be fully impregnated and large quantities of
resin flow through the distribution medium before infusion is complete. If resin
gelation occurs before the dry area is fully impregnated, a dry patch remains in the
laminate.
A good strategy must be used to complete the impregnation. As mentioned earlier,
a pressure gradient is present in the cavity hence the thickness and thus the fibre
volume fraction of the impregnated preform are not uniform. In order to optimise
the quality of the final component, fibre content must maintain the design
specification; in other words the thickness of the laminate must be controlled. This
can be achieved by subjecting the entire mould cavity to the lowest absolute
pressure present in the cavity, i.e the vacuum set at the outlet. The solution, called
bleeding, consists of closing the inlet and letting excess resin flow out of the mould
until pressure equilibrium is reached in the cavity. However, this technique presents
some risks since resin formulation must be adjusted in order for gelation not to
occur before a uniform thickness distribution is obtained. In addition it is not
possible to define the best time to stop the bleeding process. One risk is to let too
much resin flow out, which results in resin starvation in the laminate. Qi [21]
proposed a model of bleeding, focusing on the time required to let the excess
volume of resin flow out of the mould, from the end of the infusion to resin gelation.
Consequently bleed-out is highly dependent on the resin viscosity and the amount
of catalyst incorporated. It is therefore difficult to rely on gel time to end the
infusion and guarantee a good quality component with no resin starvation.
Hoebergen [22] proposed a solution to minimise that risk. The technique consists of
closing both inlet and outlet, respectively at atmospheric pressure and under
vacuum, at the end of the infusion so that the pressure gradient present in the
cavity finds equilibrium between inlet and outlet. In this case, the thickness and
fibre content are uniform and any risk of resin starvation is eliminated. However,
fibre content is not maximised.
- 10 -
Infusion time in VI is directly dependent on preform thickness, pressure gradient
and permeability variations in the preform [1, 10, 15, 23]. Thibeaudau [19] reports
the influence of preform thickness on infusion time. However, the importance of
flow layer permeability compared to that of the preform is not explained. A solution
to control lead-lag and prevent dry patches is proposed by Heider [23] who
developed a double vacuum bag. The outer bag has an embedded network of
parallel channels while the inner bag, in contact with the preform, has a smooth
surface. Air is extracted between the two bags so that the inner bag conforms to the
shape of the outer one (see Figure 1.6). The control of the vacuum between the two
bags allows channel size to be adjusted and thus flow velocity in the parallel
channels can be controlled. The system allows fast impregnation even with high
viscosity resin. Though, all of the channels on the outer bag have the same
dimensions which implies that the permeability of the channels formed on the inner
bag is constant on the entire surface of the preform.
No vacuum. The inner bag Vacuum is on. The inner bag
is a flat membrane conforms to the shape of the outer bag
Fig. 1.6. Double vacuum bag system [23].
This system is not desirable in cases where the permeability of the preform is not
uniform. The latter is due mainly to the complex shape of the component where the
fabric is sheared and compacted in certain areas such as corners and edges. Non-
uniform permeability distribution also occurs when the preform is made of different
types of fabric. Andersson [15] uses an existing analytical model to calculate an
average permeability of the preform, combining the reinforcement and the high
permeable layer. Still, results are uncertain since the thickness of each material has
to be known and cannot be measured accurately during VI. Only an average value
of maximum and minimum permeability of the overall preform can be determined
by measuring the thickness of the dry preform, under vacuum and prior to infusion,
and the average thickness of the laminate after cure. Thickness variations during VI
Outer bag
Inner bag
Reinforcement
- 11 -
were measured and models were developed in order to predict them [17, 20, 23].
Thickness changes during VI imply that the compaction of the fabric has to be
considered. Several models have been developed in order to describe and predict
flow in VI [1, 24, 25]. Hammami et al. [24, 25] proposed a one-dimensional model
for VI that takes into account the change in cavity height during infusion by
considering a succession of cases with different constant thicknesses that symbolise
thickness changes during impregnation. Permeability values used in the model were
taken from independent experimental measurements on dry fabric (i.e transient
permeability). However, this is not accurate since in VI the fabric behind the flow
front is wet and the permeability is different from that measured experimentally (i.e
steady-state permeability behind the flow front and transient permeability at the
flow front).
Research in VI has focused on experimental analysis and modelling of the process.
However, a full characterisation of the infusion technique in order to determine the
effect of key parameters involved in VI is not available. Past research focused on
the use of a highly permeable layer to enhance flow and on the interest of having
such a homogeneous structure to guarantee uniform flow enhancement in the
mould cavity. Specific studies of lead-lag in VI have been conducted by several
authors and the characterisation of the flow through the highly permeable layers
was studied experimentally. Though, the way lead-lag is created has not been
demonstrated. In addition, subjects such as the effect of thickness changes and
overlapping in the preform or the influence of complex mould geometry and fabric
bending on flow progression and lead-lag have not been investigated. Flow
behaviour through the distribution medium has seen little attention. Such analyses
are necessary to get accurate process simulations in VI. Finally this review shows
opportunities to develop new highly permeable layers where, instead of considering
the material as a homogeneous permeability distribution, the permeability of the
flow layer is adapted to that of the preform infused. This would allow each
component to have its own specific high permeable layer, optimised for minimum
infusion time with no defects.
1.4. Theme of the study
This work has been conducted in collaboration with several industrial partners in the
context of a Foresight Vehicle project (2001-2004) called Roadlite. The general aim
- 12 -
of Roadlite was to design, develop and manufacture a lightweight semi-trailer using
composite materials technology as an alternative to heavy and labour intensive
steel chassis trailers currently available on the market. The project focused on the
development of a chassis for a semi-trailer where mechanical properties of
composite materials, combined with their low weight and high design flexibility
would allow a saving of 25 % in the tare weight as well as increasing the inner
volume of the trailer. This includes the use of an innovative manufacturing process,
i.e vacuum infusion, to produce the composite chassis.
The University of Nottingham has focused on the manufacturing method. Its role
was to study the VI process so that it can be more predictable, controllable,
adjustable and repeatable, allowing industries to save time and money by reducing
trial and error prior to production.
The study of VI has been split into two main sections. One, which is reported in this
thesis, consisted of characterising VI experimentally and modelling the effects of
parameters considered to be key factors in the process. VI was analysed by
studying process and structural parameters. Lead-lag creation and evolution during
preform impregnation was investigated fully. In addition flow through the
distribution medium, responsible for lead-lag creation, was modelled using CFD
(Computational Fluid Dynamics) and analysed experimentally. Finally a design
methodology of an innovative integrated distribution medium in a vacuum
membrane was proposed and void formation through such a structure was
modelled. The other part of the VI study, published by N. C. Correia [26], presents
a mathematical model for flow in vacuum infusion that takes into account
compaction and thickness changes of the preform during impregnation.
1.5. Objectives of the study
The objectives of this study are to understand and demonstrate flow phenomena in
vacuum infusion in order to improve flow control and preform impregnation. After a
full understanding of the process the work focuses on the study of lead-lag in order
to demonstrate how it is created during vacuum infusion and to investigate which
parameters are responsible for its creation and evolution, and how it can be
controlled. This requires a thorough analysis of flow through the distribution
medium, identified to be responsible for lead-lag creation, when combined with a
- 13 -
preform of lower permeability. However, flow behaviour through this material is not
easily identifed and needs to be modelled using CFD analysis. Therefore the
presence of a distribution medium on top of the preform can be considered when
modelling flow in VI and realistic flow conditions can be used for accurate
modelling. Finally results from the analysis open possibilities for further
improvement of the VI process in order to make it more repeatable, controllable
and cost effective. This is achieved by proposing a new approach to VI where the
distribution medium is complementary to the preform and its permeability adjusted
to that of the preform. The principal objective is to adjust lead-lag in order to better
control the overall flow front progression and therefore simplify infusion strategies
and flow predictions.
- 14 -
References
1. K. Han, S. Jians, C. Zhang, B. Wang. Flow modelling and simulation of SCRIMP
for composites manufacturing. Composites Part A, 31, 2000, p. 79-86.
2. W.D. Brouwer, E. Van Herpt, M. Labordus. Vacuum injection moulding for large
structural applications. Composites Part A, 34, 2003, p. 551-558.
3. M. Koefoed, E. Lund, L. Lilleheden. Modelling and simulation of the VARTM
process for wind turbine blades. 10
th
European Conference on Composite
Materials (ECCM-10), June 3-7, 2002, Brugge, Belgium.
4. I. J. Verhaeghe. Composittrailer introducing composites in a conservative
environment. 10
th
European Conference on Composite Materials (ECCM-10),
June 3-7, 2002, Brugge, Belgium.
5. G. Weijde, A. Brodsjo, J. Tyrrell, L. Van Rijn, M. Brooker. Development of low
cost composite fairings for the ariane 5 engine thrust frame applying vacuum
assisted RTM. 23
th
International SAMPE Europe Conference, April 2002, Paris,
France, p. 493-502.
6. Vetrotex website : http : \\www.vetrotex.com
7. S. Bickerton, E.M Sozer, R.J Graham, S.G Advani. Fabric structure and mold
curvature effects on preform permeability and mold filling in the Resin Transfer
Molding process. Part 1: experiments. Composites Part A, 31, 2000, p. 423-
438.
8. K-T. Hsiao, R. Mathur, S. G. Advani, J. W. Jr Gillespie, B. K. Fink. A closed
form solution for flow during the vacuum assisted resin transfer molding
process. Journal of Manufacturing Science and Engineering, Aug 2000, 122, p.
463-475.
9. W. H Seemann. Plastic transfer moulding techniques for the production of fibre
reinforced plastic structures. US patent No 4902215, filed March 1989.
- 15 -
10. T. Searle, J. Spooner, S. Grove, R. Cullen, S. Davy. Manufacture of large
composite structures by vacuum resin infusion. 24
th
International SAMPE
Europe Conference, April 2003, Paris, France, p. 113-120.
11. M.K. Kang, W.I. Lee, H.T. Hahn. Analysis of vacuum bag resin transfer
moulding process. Composites Part A, 32, 2001, p. 1553-1560.
12. M. Feiler, L. Ischtschuk. Vacuum assisted resin infusion (VARI): on the way to
serial production. 24
th
International SAMPE Europe Conference, April 2003,
Paris, France, p. 683-691.
13. C. Williams, J. Summerscales, S. Grove. Resin infusion under flexible tooling
(RIFT): a review. Composites Part A, 27, 1996, p. 517-524.
14. X. Sun, S. Li, L.J. Lee. Mold filling analysis in vacuum assisted resin transfer
molding. Part I: SCRIMP based on a high permeable medium. SPE Technical
journals, Polymer Composites, Dec. 1998, 19, N6, p. 807-817.
15. M. Andersson, S. Lundstrom, B.R. Gebart, R. Langstrom. Development of
guidelines for the vacuum infusion process. 8
th
Fibre Reinforced Composite
Conference (FRC), Sep 13-15, 2000, Newcastle, UK, p. 113-120.
16. H. M. Andersson, T.S. Lundstrom, B.R. Gebart, R. Langstrom. Flow enhancing
layers in the vacuum infusion process. Polymer composites, Oct 2002, 23, N
o
5,
p. 895-901.
17. A. Hammami, B.R. Gebart. Experimental investigation of the vacuum infusion
molding process. 8
th
European Conference on Composite Materials (ECCM-8),
June 1998, Naples, Italy, p. 577-584.
18. J. Ni, S. Li, X. Sun, L.J. Lee. Mold filling analysis in vacuum assisted resin
transfer molding. Part II: SCRIMP based on grooves. Polymer Composites,
Dec. 1998, 19, N6, p. 818-829.
- 16 -
19. M. Thibaudeau, R.A. Shenoi. Understanding flow behaviour in the resin infusion
process. 10
th
European Conference on Composite Materials (ECCM-10), June 3-
7, 2002, Brugge, Belgium.
20. X. Song, A.C. Loos, B. Grimsley, R. Cano, P. Hubert. Modelling manufacture of
textile composites by VARTM. 6
th
International Conference on Textile
Composites (TexComp 6), Sept 2002, Philadelphia, USA.
21. B. Qi, J. Raju, T. Kruckenberg, R. Stanning. A resin film infusion process for
manufacture of advanced composite structures. Composite Structures, 47,
1999, p. 471-476.
22. A. Hoebergen, J. Holmberg. Vacuum infusion, ASM Handbook, 21, Composites,
Materials Park (OH), 2001, p. 501-515.
23. D. Heider, A. Karamitsos, J.W Gillespie, S. Walsh, E. Rigas, W. Spurgeon, W.
Roy. Experimental validation and optimisation of the FASTRAC process. 22
th
International SAMPE Europe Conference, April 2001, Paris, France, p. 293-303.
24. A. Hammami, B.R. Gebart. A model for the vacuum infusion molding process.
7
th
Fibre Reinforced Composite Conference (FRC), 1998, Newcastle, UK, p.
136-145.
25. A. Hammami, B.R. Gebart. Model for vacuum infusion molding process.
Plastics, Rubber and Composites Processing and Application, 1998, 17, N4.
26. N. C. Correia. Analysis of the vacuum infusion moulding process. PhD thesis,
2004, University of Nottingham.
- 17 -
Chapter 2
General analysis of the process and structural parameters
2.1. Introduction
In VI, flow relates directly to process and structural parameters which control the
impregnation of the preform. This chapter aims to characterise experimentally flow
phenomena in vacuum infusion in order to understand the influence of key process
parameters on preform impregnation and thickness distribution. The parameters
studied here are process variables, such as bagging materials and vacuum level,
and materials variables i.e resin and reinforcement. They were varied systematically
during VI experiments and their effects on process time and preform compressibility
- relaxation response were measured. Four experimental studies were conducted.
One consisted of measuring the effect of consumables (breather, bleed out fabric,
vacuum membrane, distribution medium) on resulting infusion times. The second
study focused on the effect of fabric style on infusion times and the corresponding
compaction and permeability-porosity relationships. The control of resin supply at
the end of the infusion, i.e bleeding technique, was also analysed to measure its
effect on thickness distribution.
2.2. Background
In the conventional VI process a stiff mould half is covered with a stack that
includes a reinforcement, a bleed out fabric, a resin distribution medium and a
breather felt (see Figure 1.1). A flexible membrane seals the cavity. An inlet is
connected to an atmospheric supply of liquid resin. The outlet is normally connected
to a vacuum pump, via a resin trap. The cavity is evacuated and resin is admitted.
Flow is driven by the pressure difference and when the cavity is judged to be filled,
the resin supply is usually isolated. The part is debagged only after resin cure.
Feiler [1] mentions that VI is a cost effective, reliable, robust processs, giving high
and repeatable part quality. He reports, however, a lack of development in quality
control of the lay-up and vacuum bagging.
- 18 -
Some technical studies related to VI have been reported. Andersson [2] developed
process guidelines based upon the evolution of preform thickness during infusion
and the out-of-plane flow front shape created by the use of a highly permeable
continuous strand mat on top of the preform. Flow enhancing layers produced large
reductions in infusion times and a steady out-of-plane flow pattern named as lead-
lag. Thickness measurements confirmed the time-dependent transverse compliance
of the laminate, compaction was observed after the resin inlet was closed and the
cavity pressure became hydrostatic. Feiler [1] mentions that the impregnation of
the fabric is achieved through the thickness of the preform and that the fibre
content of the laminate is controlled by the pressure gradient in the cavity as well
as resin viscosity.
The physics of this type of infusion process are captured adequately by Darcys law
and this provides a useful indication of the important variables:
dx
dP K
v

=
u c
2.1
where v is the interstitial velocity of the fluid, observed at microscopic scale and
corresponds to the speed of the fluid particles.
Another type of velocity called superficial velocity (macroscopic scale) is also
employed. A relationship exists between interstitial velocity and superficial velocity
as described in Equation 2.2.
c
u
v = 2.2
where u is the superficial velocity of the fluid (m/sec).
The interstitial velocity v corresponds to the observed flow rate. Another expression
of v is defined as:
c
=
S
V
v 2.3
- 19 -
The superficial velocity u corresponds to the flow rate through the section S and
does not depend on the porosity.
During infusion, the pressure field becomes modified from that occurring during the
simpler case of RTM because the domain of interest no longer involves a uniform
thickness, porosity or permeability. For rectilinear flow this results in a non-linear
pressure distribution as shown in Figure 2.1 and efforts are now under way to
simulate this effect [3], taking into account the thickness changes that occur during
the impregnation of the preform.
Fig. 2.1. Pressure distribution obtained in RTM and VI [8].
Unlike a fixed cavity process, such as RTM, the permeability of the fibre bed is
dependent upon its transverse (thickness direction) compliance. This is because one
face of the preform is subjected constantly to a pressure boundary condition.
Various empirical compaction models can be used to represent this effect. Perhaps
the simplest is that adopted in the review by Robitaille [4]:
B
f
AP V = 2.4
Where P is the compaction pressure, A the fibre volume fraction at a compaction
pressure of 1 Pa and B is the compaction stiffening index.
0
20
40
60
80
100
0.0 0.2 0.4 0.6 0.8 1.0
Distance between inlet and oulet (m)
P
r
e
s
s
u
r
e
(
k
P
a
)
VI
RTM
- 20 -
Hammami studied transverse compliance [5, 6, 7] and proposed a one-dimensional
model [5] of the process, relating it to changes in cavity thickness occurring during
the infusion. A power law was used to describe the compressive behaviour of the
preform. The largest changes in preform thickness occurred at the beginning of the
infusion when the flow rate was greatest. Also, the flow front velocity decreased as
the impregnated length increased and that was affected by resin viscosity. It was
concluded that Darcys law can be used to describe VI if thickness changes are
considered.
Hammami [7] also studied thickness variations of the preform during VI. These
were recorded before infusion, on dry fabric and at different stages of the
impregnation. Results showed that preform impregnation resulted in thickness
variations, i.e thickness increased compared to the thickness of the dry preform,
with the largest variations occuring near the inlet. At any section of the preform this
variation decreased as the flow front progressed and reached a constant level. The
latter was assumed to be related to a critical infusion length and preform
permeability.
The author highlights that stiff reinforcements are less sensitive to compressibility -
relaxation during preform impregnation and existing compaction models used to
describe compaction behaviour of the preform are not accurate since they
underestimate compressive response of the reinforcement at low pressures.
Williams [9] studied experimentally fabric compression and the effect of
reinforcement and resin flow on final component thickness. The author explains that
in VI there is no direct control of the thickness of the component and that it
depends on the compressibility - relaxation of the material under a given pressure
as well as interactions with bagging materials. The compression of the
reinforcement is also influenced by lubrication of the material during infusion. In
addition because of the evolution of the pressure gradient the compressive force
applied to the preform is not constant. Experiments highlight thickness changes in
the preform during VI. Results show that prior to infusion the thickness of the dry
preform decreases due to the compaction of the fabric under vacuum. The
lubrication of the reinforcement due to the flow of the resin compacts the preform
further. After resin has passed a given point, the thickness of the preform at this
point increases due to a decrease of the pressure gradient behind the flow front.
- 21 -
The same authors also studied parameters involved in VI and their effect on the
process. Infusion strategy, preform relaxation and resin temperature dominated.
Resin temperature can be particularly critical due to the steep nature of the
viscosity-temperature curve around room temperature for most candidate resins.
Authors report different methods to observe and record flow front progression
through the preform thickness during VI. Thibaudeau [10] used the Smartweave
recording system. Single skin preforms (single side infused) and sandwich
structures (double side infused) were processed. In single skin laminates the
authors studied the thickness, shape and size of the preform as well as the type of
reinforcement. A layer of highly permeable medium was used on top of the preform
but the type and permeability of the material is not mentioned. The authors report
the creation of a lead-lag due to the occurrence of a trans-plane flow. They also
mention the creation of micro-flows that infuse the lower side of the preform ahead
of the main flow front, creating radial flows. These micro-flows are then absorbed by
the main flow. They conclude that the time to fully impregnate the preform is
directly dependent on preform thickness and so number of reinforcement layers. It
is also mentioned that the velocity of the flow converges to the same value, but not
at the same time, whatever the thickness of the preform. The authors report that
the type of reinforcement modifies trans-plane flow but no more information is
given and no permeability data are available. It is only mentioned that micro-flow
appears in some of the reinforcements due to their structure. Another technique to
record flow front progression through the thickness of the preform was that
proposed by Andersson [2] who used a visual method to record trans-plane flow in
VI. It consisted of using colour marks on the different layers of the preform. The
colour contours then appeared successively with the impregnation of the layers.
This chapter aims to obtain an overview of the vacuum infusion process, from lay-
up of the preform to infusion and laminate thickness distribution. The objective is to
identify process and structural parameters that influence VI the most especially in
term of infusion time, preform impregnation and laminate thickness. Process
parameters are bagging materials, vacuum level and infusion strategy (position of
inlets and outlets, opening - closing of inlets and outlets at the start and termination
of the infusion). Structural parameters involve the preform and the fluid.
- 22 -
y = 13324x
-1.2996
0
50
100
150
200
250
300
350
400
15 25 35 45 55 65 75
Temperature (
o
C)
V
i
s
c
o
s
i
t
y
(
m
P
a
.
s
)
2.3. Experimental analysis: materials
2.3.1. Reinforcement
Three E-glass reinforcements were chosen to study the effect of fibre architecture
on the process and resulting laminate thickness: a triaxal fabric FGE 117, 1173 g.m
2
from Formax, a plain weave RT600, 600 g.m
2
and a continuous filament random
mat U751 / 375, 375 g.m
2
both from Vetrotex. Technical characteristics of the
fabrics are available in Appendix 2.
2.3.2.Fluid
Two types of experiments were conducted. The first involved a mineral oil HDX 30
from Trent Oils. The other involved a pre-accelerated polyester resin 701 PAX from
Scott Bader catalysed with 1 wt% of Butanox M50 from Akzo Chemie. Technical
characteristics of the fluids are described in Appendix 2.
The viscosity of the oil and uncatalysed resin was measured and results are
displayed in Figures 2.2 and 2.3. The aim was, first to measure what would be the
effect of room temperature variations on oil and resin viscosity, second to compare
the viscosity of both the oil and the resin at different temperatures.
Fig. 2.2. Viscosity of HDX 30 oil as a function of temperature. Power curve fit.
- 23 -
y = 20179x
-1.5338
0
20
40
60
80
100
120
140
160
180
20 30 40 50 60 70
Temperature (
o
C)
V
i
s
c
o
s
i
t
y
(
m
P
a
.
s
)
Fig. 2.3. Viscosity of uncatalysed 701 PAX polyester resin
as a function of temperature. Power curve fit.
In addition the evolution of resin viscosity and temperature as a function of various
initiator loadings was measured and results are displayed in Appendix 3. The
objective was to measure the time available between resin - catalyst mixing and the
occurrence of gelation. During VI experiments, this would give an indication on the
time available to infuse a preform while the resin mix displays a constant viscosity.
All measurements were conducted on a Brookfield viscometer LV-DII.
2.3.3.Consumables
A woven polyamide bleed out fabric 60 B from Tygavac was used with an HDPE,
chained stitch distribution medium H8036 from Palmhive Textiles (100 g/m
2
). A
non-woven polyester breather NW85 from Tygavac (85 g/m
2
), an impermeable
Mylar

film, to prevent impregnation of the breather fabric, and a polyamide


vacuum film either textured with an approximately 0.3 mm deep cracked ice
pattern or 0.05 mm thick smooth surface were also added. The assembly was
sealed by a strip of 12 mm wide DBT 101 mastic tape from Tygavac. Technical
characteristics of the consumables are described in Appendix 2.
- 24 -
2.4. Assumptions
2.4.1. Bagging strategy
The use of a textured vacuum bag in VI aims to reduce infusion times thanks to the
random channels present on its surface, which create preferential paths for the resin
to flow. Therefore it is assumed that this material can replace a distribution medium
in term of flow enhancement properties. In addition because of its greater stiffness
compared to a smooth bag, the textured bag is assumed to be less deformable
under vacuum thus increasing further porosity underneath the bag. However this
should result in a lower conformation of the preform under vacuum and decrease
the fibre volume content in the laminate. It is assumed that the distribution medium
does not affect the compaction of the preform.
The flow enhancement properties of the textured bag should be masked when a
breather felt is employed since the latter prevents direct contact between the bag
and the preform. In addition the low conformation of the bag under vacuum should
reduce or eliminate the properties of the breather felt, which seeks to maximise air
extraction in the mould cavity and thus increases preform compaction. In the case a
smooth bag is used, the additional compaction effect of the breather felt should
reduce or totally mask the flow enhancement properties of the distribution medium.
Still a breather felt is expected to reduce resin consumption and improve preform
compaction, thus increase fibre volume content in the laminate.
Infusion time in VI is assumed to highly depend on the permeability - porosity
relationship between the high permeability layer and the preform. In the present
work a polymer mesh was used as high permeability layer but some authors report
the use of a chopped strand mat as a flow enhancing layer on top of a non-crimp
stitch bonded reinforcement [2]. Leenders [10] used a roving or a continuous
filament random mat (CFRM) as a highly permeable layer on top of a preform made
of roving and unidirectional glass layers.
- 25 -
2.4.2. Fabric style
The continuous filament random mat is a highly porous and thick material compared
to plain weave and triaxial fabrics. Therefore CFRM laminates should be infused
quickly but present low fibre content.
The plain weave is expected to have a much lower permeability than the CFRM,
closer to that of the triaxial. In addition its propensity to compaction is expected to
be much lower than the CFRM. However the material is very thin and for a same
number of plies should result in a much thinner preform than the CFRM or Triaxial.
Infusion times are expected to be high due to the low permeability of the material
(see Appendix 1). But, due to the use of a distribution medium layer, trans-plane
flow will occur which should significantly decrease infusion time due to the short
distance to travel through the thickness of the preform.
Finally the triaxial fabric is expected to have a much lower permeability than the
CFRM. In addition the material is thick and expected to have low compressibility -
relaxation behaviour. Consequently infusion times are assumed to be longer than
with the other preforms even with the use of a distribution medium. However the
laminate will present higher fibre content, i.e lower resin consumption.
Due to the use of a distribution medium on top of the preforms, the latter are
expected to show a lead-lag during infusion. Its length will depend on the
permeability of each material. The CFRM is assumed to have a permeability close to
that of the distribution medium and is expected to show a very short lead-lag. The
plain weave, due to its low thickness and the occurrence of a trans-plane flow, is
expected to show a short lead-lag although the material presents a low
permeability. Finally the triaxial fabric is expected to present a large lead-lag, this
material being both thick and not highly permeable.
2.4.3. End of infusion
The aim is to measure the effect of bleeding on preform compaction and compare its
efficiency with different types of preforms, i.e high and low compressibility -
relaxation types.
- 26 -
Due to its very porous structure the CFRM is assumed to present a high
compressibility - relaxation behaviour contrary to plain weave and triaxial fabrics.
Therefore bleeding should have more effect on CFRM. Yet the time required to reach
a hydrostatic pressure inside the cavity is not predictable neither is the thickness
variation of the preform.
2.5. Experimental analysis: results
2.5.1. Bagging strategy
Moulding procedure
Only continuous random filament mat U 751 / 375 was involved in this series of
tests. The experiments consisted of studying the effect of bagging sequence on flow
properties. Table 2.1 describes the test matrix for the experiments conducted. All
preforms were infused with mineral oil.
Table 2.1. Test matrix: effect of bagging strategy on flow properties in VI
Test number 1 2 3 4 5 6
Fluid Oil HDX 30
Reinforcement U 751 / 375
Absolute pressure inside bag (mbar) 101.3
Number of plies 9
Bleed out fabric (BOF)
Distribution medium (DM) - -
Breather felt - - - - Bagging materials
Vacuum bag
T = Textured
S = Smooth
T T T S S S
The reinforcement was placed on a 300 mm by 500 mm Perspex plate as shown in
Figure 2.4. The stacking sequences are given in Table 2.1. The bag was sealed and
checked for air leakage. The inlet was connected to a vented container containing
the oil at room temperature. The outlet was connected, via a resin trap, to an
Edwards RV5 pump. Both inlet and outlet used 6 mm bore PVC pipe connections
with 15 mm diameter steel springs to form full width inlet and outlet galleries. The
objective was to promote rectilinear flow throughout the cavity.
Flow front progression was monitored by video camera on the lower face and
visually analysed on the upper face. Times at which the flow front reached points
- 27 -
located at 50 mm spacings along the centreline of the preform, on both upper and
lower face of the mould, were recorded (0 mm = position of the inlet, 500 mm
position of the outlet).
The viscosity of the oil was recorded before each experiment on a Brookfield LV-DII
in order to measure the effect of viscosity variations on infusion times. Results are
available in Appendix 3.
Fig. 2.4. Experimental set-up: effect of bagging strategy on flow properties in VI.
They show that all recorded viscosity values were comprised between 200 mPa.s
and 270 mPa.s. It represents a variation of 7.5 % of the average viscosity value
obtained from all these measurements. It was assumed that the influence of the oil
viscosity on the interpretation of the time recordings was negligible.
Results
All experiments used a 9layer preform of continuous filament random mat U 751 /
375 terminated by a layer of bleed out fabric. The main variables were the
introduction of a distribution medium, a breather felt and the use of non-textured
versus textured bagging films. The latter was used to compare its flow enhancement
properties with that of the distribution medium. The infusion tests were all
conducted with an absolute pressure at the vent of 101.3 mbar.
Figures 2.5 and 2.6 show the infusion time for the first 500 mm of flow front
progression and for the different tests conducted.
Inlet
Outlet
Video
camera
Mirror
Preform and
bagging
materials
Wet preform
Flow front Dry
preform
- 28 -
Fig. 2.5. Evolution of flow front position with time.
The shortest infusion time involves a textured bag and a distribution medium. The use of a
breather felt reduces the flow enhancement properties of the distribution medium by extracting
more air from the cavity, thus compacting the preform further.
Fig. 2.6. Evolution of flow front position with time.
The shortest infusion time involves a distribution medium. The breather felt compacts the
preform further, masks the flow enhancement properties of the distribution medium and results
in longer infusion time.
The strongest effect here appears to be that of the vacuum bag and the textured
bag shows generally shorter infusion times than the smooth film. As expected it
0
20
40
60
80
100
120
140
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
)
Textured / BOF
Textured / BOF / DM
Textured / BOF / DM / Breather
0
20
40
60
80
100
120
140
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
)
Smooth / BOF
Smooth / BOF / DM
Smooth / BOF / DM / Breather
- 29 -
enhances flow because of the additional porosity introduced. However the use of a
smooth bag combined with a distribution medium shows infusion times very close to
the textured bag and the textured bag - distribution medium strategies, respectively
slightly lower and higher.
This tends to show that, first the textured bag presents similar flow enhancement
properties to the distribution medium, second that the combination textured bag -
distribution medium does not further reduce infusion time significantly.
The textured bag is thicker than the smoooth bag, 0.3 mm against 0.05 mm. The
greater stiffness provided by the extra thickness might be expected to reduce
conformance of the preform, thereby increasing porosity further. Recordings of fluid
consumption during preform impregnation confirm this hypothesis as the quantity of
oil infused is higher when using the textured bag, as shown in Figures 2.7 and 2.8
(during the experiments, the oil container was placed on a digital scale and mass
readings were recorded when the flow front reached 50 mm marks on the centreline
of the lower side of the mould).
Fig. 2.7. Evolution of fluid consumption with flow front position.
As expected the distribution medium increases fluid consumption. The extra compaction
provided by the breather felt is seen through a reduction of fluid consumption.
We assume that, for a similar level of vacuum, the laminate infused under the
textured bag presents a lower fibre content than the laminate infused under the
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0 100 200 300 400 500 600
Flow front position (mm)
O
i
l
(
c
m
3
)
Textured / BOF
Textured / BOF / DM
Textured / BOF / DM / Breather
- 30 -
smoooth bag - distribution medium combination. These results show that the short
infusion times obtained with the textured bag are more due to its low conformation
under vacuum than its effective flow enhancement properties. Therefore smooth
bag - distribution medium strategy is preferred.
Fig. 2.8. Evolution of fluid consumption with flow front position.
The use of a distribution medium increases fluid consumption. The further compaction provided
by the breather felt is greater than with the textured bag (lower fluid consumption) due to a better
comformation of the smooth bag under vacuum. Generally smooth bag strategies present a
lower fluid consumption than textured bag strategies.
Adding a breather felt does not reduce infusion time further, rather it tends to
increase it. Indeed the breather compacts the preform further which tends to reduce
both the flow enhancement properties of the distribution medium and the
permeability of the reinforcement. This is confirmed by the fact that fluid
consumption is reduced when a breather felt is added to the stack. Therefore the
use of such a material in VI is not a prerequisite although it contributes to increase
the fibre content in the laminate.
Finally the effect of the distribution medium on infusion time is visible but not very
significant in the present cases. This is attributed to the high permeability of the
continuous filament random mat being close to that of the distribution medium.
Though, the high porosity of the distribution medium is seen through the higher
quantity of fluid infused in the mould cavity when this material is used in the
bagging strategy.
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0 100 200 300 400 500 600
Flow front position (mm)
O
i
l
(
c
m
3
)
Smooth / BOF
Smooth / BOF / DM
Smooth / BOF / DM / Breather
- 31 -
2.5.2. Effect of fabric style
Moulding procedure
These experiments involved all reinforcements. The bagging sequence consisted of a
bleed out fabric, distribution medium and smooth vacuum bag. The vacuum level
was of 101.3 mbar of absolute pressure. The study consisted of investigating the
effect of reinforcement structure and permeability on preform impregnation and
infusion time. Table 2.2 shows the test parameters used.
300 mm by 500 mm 9-layer preforms were placed on a 1000 mm by 1000 mm
glass plate as shown in Figure 2.9. Consumables were added, the bag was sealed
and checked for air leakage. The inlet was connected to a vented container
containing the resin at room temperature.
Table 2.2. Test matrix: effect of reinforcement type on flow and laminate properties
Test 5 7 8
Fluid Resin 701 PAX
Reinforcement U 751 / 375 RT 600 FGE 117
Absolute pressure (mbar) 101.3 101.3 101.3
Number of plies 9
Bleed out fabric
Distribution medium Bagging materials
Smooth vacuum bag
The outlet was connected, via a resin trap, to an Edwards RV5 pump. As for
experiments involving oil, both inlet and outlet used 6 mm bore PVC pipe
connections with 15 mm diameter steel springs to form full width inlet and outlet
galleries in order to promote rectilinear flow throughout the cavity. Additions were
made at room temperature and blended using an air mixer at 1500 RPM for 2
minutes. Entrapped air was expelled using a vacuum chamber under 200 mbar of
absolute pressure for 10 minutes. Flow front progression was monitored by video
camera on the lower face of the mould.
In addition flow through the thickness of the preform was monitored using the
SmartWeave system. This technique uses a grid made of conductive fibres to
detect the resin flow front position. The grid, made of sensing lines and excitation
lines, is laid-up in the preform. Times at which the flow front reached points located
at 50 mm spacings along the centreline of the preform, on both upper and lower
- 32 -
face of the mould, were recorded (0 mm = position of the inlet, 500 mm position of
the outlet).
Fig. 2.9. Experimental set-up: effect of fabric style on flow properties in VI.
A continuous measurement of the DC resistance at each intersecting node is
obtained through an electronic package and a LabVIEW software program. The
excitation lines generate a DC voltage while the sensing lines detect the voltage.
The arrival of resin at each intersecting node is detected through an increase of the
DC resistance as shown in Figure 2.10. The grid can be made of any conductive
filaments but, since it is permanently cured into the laminate, preferably of the
same nature as the preform.
In the experimental analysis presented here, the conductive grid was made of
carbon fibre roving. Eight by five filaments grid was used to record flow front
position through the thickness of the preform. The excitation lines were placed on
the third layer of the preform while the sensing lines were laid on the sixth layer of
the preform. Experiments with triaxial fabric involved two grids of eight by five
filaments. The first grid was placed on the first and third layer of fabric respectively
while the second grid was laid on the sixth and eighth layer.
Resin trap
Pump
Spring (outlet)
Glass plate
Mirror
Spring (inlet)
Preform and bagging materials
- 33 -
Fig. 2.10. Typical recordings with SmartWeave system.
Results
Figure 2.11 compares flow front progressions in the continuous filament random
mat U 751 / 375, the triaxial FGE 117 and the plain weave fabric RT 600. The plain
weave and continuous filament random mat produced short infusion times, despite
major differences in porosity at the nominal compaction pressure of 900 mbar (see
Figure 2.12).
Conventionally, the lower porosity in the plain weave would be expected to reduce
permeability and produce a low fluid velocity but this was not the case low
porosity also means that less resin is required to fill the mould cavity, as shown in
Figure 2.13. The plain weave presents the lowest fluid consumption although it has
higher porosity than the triaxial fabric. This is due to the much lower thickness of
the plain weave plies which results in a thinner preform than the triaxial for an
identical number of plies.
Inlet
Outlet
Resin flowprogression
Resin flow progression, distance from inlet
Excitation line
Voltage at node
intersection as a
function of time
Intersecting node
Dry preform
Wet preform
Sensitive line
t
1
t
2
t
3
Resin detection at node intersection
- 34 -
Fig. 2.11. Evolution of flow front position with time.
CFRM, of permeability close to that of the distribution medium, presents no lead-lag while triaxial
presents the largest lead-lag due to both low permeability and high preform thickness. The plain
weave preform presents a short infusion time and short lead-lag, although its permeability is low,
due to a very low thickness.
Fig. 2.12. Evolution of fibre volume fraction with compaction
pressure for the different reinforcements tested [8].
As expected material compaction shows higher values for saturated preforms. At 9.10
4
Pa of
compaction pressure, the CFRM and plain weave represent respectively an average of 38% and
86% of the transient triaxial fibre content. In the saturated case, the fibre volume fraction of the
triaxial is respectively 60% and 16% higher than that of the CFRM and the plain weave. The
plain weave shows the lowest variation in fibre content between transient and saturated state, i.e
1.3 % of increase from transient to saturated at 9.10
4
Pa compared to 4.8% for the triaxial and
11% for the CFRM at the same compaction pressure.
0
100
200
300
400
500
600
700
800
900
1000
0 100 200 300 400 500
Flow front position (mm)
T
i
m
e
(
s
)
U 751 / 375
RT 600 upper
FGE 117 upper
RT 600 lower
FGE 117 lower
0
10000
20000
30000
40000
50000
60000
70000
80000
90000
100000
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70
Fibre volume fraction
C
o
m
p
a
c
t
i
o
n
p
r
e
s
s
u
r
e
(
P
a
)
U 751/375 Transient
U 751/375 Saturated
RT600 Transient
RT600 Saturated
FGE117 Transient
FGE117 Saturated
- 35 -
Fig. 2.13. Evolution of fluid consumption with flow front position.
Due to its low thickness and permeability, the plain weave shows low fluid consumption. The
triaxial presents a higher consumption due to a thicker preform although its permeability is low.
The CFRM preform presents the highest fluid consumption due to both high porosity and
thickness.
The triaxial filled at a markedly lower rate. The dominant feature here is thought to
be transverse compliance. The architecture of the triaxial fabric results in a
relatively high fibre volume fraction (53%) compared to the continuous filament
random mat (17%) at an equivalent wet compaction pressure.
Thus there is a large difference in permeabilities and infusion times (see Appendix
1). Clearly though, the dependence is not simply one of compliance, since the
subsequent porosity-permeability relation has also to be taken into account.
It should also be pointed out that it is misleading to explain these results simply in
terms of the in-plane permeability for the reinforcement stack alone. Even allowing
for a plug flow approximation (see Equation 2.5) [18] then the in-plane effects are
distorted by the presence of the flow enhancement medium. Figure 2.14 compares
experimental infusion times with an integration form of Darcys law (see Equation
1.2). Theoretical results reproduce the difference in infusion time due to differences
in in-plane permeability. However, theoretical infusion times are much longer than
experimental ones especially for the plain weave and triaxial fabric. This highlights
the non-representation of trans-plane permeability in the average plug-flow model
and thus only in-plane flow characterises the infusion. Because the permeability of
0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0 100 200 300 400 500
Flow front position (mm)
R
e
s
i
n
(
c
m
3
)
U 751 / 375
RT 600
FGE 117
- 36 -
the CFRM is close to that of the distribution medium, no trans-plane flow occurs and
theoretical infusion time is similar to experimental one.

=
=
n
l
l
ij
l
ij
K h
H
K
1
1
2.5
Where H is the total thickness of the preform, h is the thickness of each material, K
ij
is the permeability of each material, n the number of material layers, l the number
of different materials.
Fig. 2.14. Evolution of flow front position with time.
Comparison with an integration form of Darcys law. Theory shows much longer infusion times
than experiments for triaxial and plain weave reinforcements due to the non-consideration of
trans-plane flow in the average permeability model. Theroretical and experimental infusion times
in CFRM are similar which proves the similar permeability between the material and the
distribution medium and the non-existence of a lead-lag in the preform infused.
Figure 2.11 clearly shows the influence of the permeability difference between the
fabric and the distribution medium on infusion time. The larger the difference in
permeability the larger the difference in infusion time between the upper and lower
face of the preform. This difference is known as lead-lag and represents the
difference in time or distance between flow front progression on the upper and
lower side of the preform (see Chapter 1). This lead-lag is created by the presence
of plies of different permeability in the stack, i.e reinforcement and distribution
medium. Figure 2.15 shows the results from flow front monitoring using the
0
500
1000
1500
2000
2500
3000
3500
4000
4500
0 50 100 150 200 250 300 350 400 450
Flow front position (mm)
T
i
m
e
(
s
e
c
)
U 751 / 375
RT 600 upper
FGE 117 upper
RT 600 lower
FGE 117 lower
Polynomial (Darcy transient U 751 / 375)
Polynomial (Darcy transient RT600)
Polynomial (Darcy transient FGE117 ) )
(CFRM)
(Plain Weave)
(Triaxial)
- 37 -
SmartWeave system. Maps of infusion time as a function of flow front position
were made using visual recordings on both the upper and lower sides of the preform
as well as SmartWeave recordings through the thickness of the preform. Time is
represented with contour levels in a range of colours from blue for short times to
red for long times. Results show that no lead-lag exists in the continuous filament
random mat.
CFRM 751 / 375 No lead-lag
Results are presented in 3D with X, Y and Z
axes corresponding respectively to distance
from inlet, width of the preform and thickness
of the preform. The number of contour levels
is identical from a fabric to another but
minimum and maximum contour levels are
specific to individual fabric. Blue
represents short times and red long times.
Lead-lag is represented by a black line.
RT 600 Small lead-lag FGE 117 Large lead-lag
Fig. 2.15. Comparison of lead-lag between different types of preform.
Flow front monitoring recordings.
Due to its permeability close to that of the distribution medium, the CFRM preform does not
present any lead-lag. As expected the triaxial preform presents a large lead-lag due to a low
permeability and high thickness, which slows down trans-plane flow in the preform. Due to its
low permeability the plain weave would be expected to present a large lead-lag. However the
very low thickness of the preform results in quick trans-plane flow, thus reducing lead-lag.
Due to its high permeability, which is closer to the distribution medium permeability
than the other reinforcements, the preform does not show time difference between
the upper and lower face of the preform as shown in Figure 2.11. On Figure 2.15
lead-lag is evident for plain weave and triaxial fabrics. It is clearly much longer in
the triaxial fabric but generally its shape is not clearly defined.
0 20 s
180 190 s
0 500 s
800 1000 s
0 50 s
200 300 s
- 38 -
2.5.3. End of infusion
Moulding procedure
Techniques for ending the infusion were investigated. They consisted of either
adjusting inlet opening and closing or varying vacuum level after full impregnation
of the preform. The objective was to measure the influence of each strategy on
laminate thickness. The aim was to get a uniform thickness distribution with
maximum fibre content. The laminates were bled of excess resin until pressure
equilibrium was reached in the mould cavity. Table 2.3 shows the different tests
conducted. All tests were performed with polyester resin. The experimental set-up is
described in Figure 2.9.
Table 2.3. Test matrix: end of infusion
Fluid Resin 701 PAX
Reinforcement U 751 / 375 RT 600 FGE 117
Absolute pressure: during infusion (mbar) 101.3 101.3 101.3
20 0
0
10 10 10
Duration inlet closed before gel (min)
2 2 2
Number of plies 9
Bleed out fabric
Distribution medium Bagging materials
Smooth vacuum bag
In addition the effect of vacuum level on infusion time was investigated.
Experiments were conducted on continuous filament random mat only and preforms
were infused with mineral oil. The experimental set-up is described in Figure 2.4
and Table 2.4 shows the different tests conducted.
Table 2.4. Test matrix: effect of vacuum level on flow properties in VI
Fluid Oil HDX 30
Reinforcement U 751 / 375
Absolute pressure inside bag (mbar)
101.3
401.3
701.3
Number of plies 9
Bagging materials
Bleed out fabric
Distribution medium
Smooth vacuum bag
Flow front progression was monitored by video camera on the lower face and
visually analysed on the upper face. Times at which the flow front reached points
- 39 -
located at 50 mm spacings along the centreline of the preform, on both upper and
lower face of the mould, were recorded (0 mm = position of the inlet, 500 mm
position of the outlet).
Results
When the infusion is complete a pressure gradient is present in the cavity, resulting
in non-homogenous thickness of the laminate, which is typically higher near the
inlet, where the pressure is atmospheric.
If this pressure gradient was maintained (by leaving both inlet and outlet opened
until resin gel) then the resulting variations in fibre volume fraction and thickness
would carry over into the final composite structure. Bleeding, as studied by Qi [12]
and Hoebergen [13], seeks to extract the excess resin this is essentially a
secondary phase of vacuum consolidation. However, there is an attendant danger of
resin starvation, which can lead to voidage. As discussed by Hoebergen [13]
different strategies can be used to terminate an infusion:
- One solution consists of closing the inlet, which decreases cavity pressure and
raises fibre fraction but involves resin waste.
- If the outlet only is closed then the cavity pressure increases to atmospheric
pressure, which reduces fibre fraction.
- Another solution consists of closing both inlet and outlet at the same time. In
that case an intermediate pressure equilibrium is obtained between inlet and
outlet pressure.
- Another option consists of increasing first the pressure at the outlet and then
closing the inlet, so that in case leaks are present in the preform they do not
create any voids.
- An optimal solution consists of closing the inlet before the mould cavity is
completely filled. The priorities are to minimise void fraction and maximise fibre
volume fraction while minimising resin waste.
The present bleeding studies were made by simply closing the resin inlet and
maintaining suction at the vent for a predetermined period. The flow rate decayed
as the preform consolidated and a hydrostatic pressure was reached. The results
were compared to other experiments where the inlet had been left open until resin
gelation. Table 2.5 describes the strategy used to end each infusion.
- 40 -
Table 2.5. Details of strategy used to end infusions.
Reinforcement Duration inlet closed
before gel (min)
Gel time (min) Initial resin viscosity
(mPa.s)
20 45 284
U 751 / 375 10 43 241
2 42 261
10 45 259
FGE 117 2 44 290
0 44 266
10 55 311
RT 600 2 48 281
0 58 360
Figure 2.16 shows the corresponding thickness profiles of the laminates (3D
thickness maps of each laminate are available in Appendix 4). Thickness was
measured with a CMM (Coordinate Measuring Machine) and recorded on 45
different locations on the surface of each laminate.
In the continuous filament random mat, the laminate thickness is visibly lower when
the inlet has been closed for a longer time (see Figure 2.16). However, even after
having closed the inlet for 20 minutes, the thickness is still not uniformly
distributed. However this type of fabric presents a high compressibility - relaxation
behaviour as demonstrated in Figure 2.17, which shows the influence of vacuum
level, characterised by the absolute pressure at the vent, on flow front progression
for the continuous filament random mat U 751 / 375. Here the expected effect of
decreasing infusion times as the driving pressure gradient increases is seen. The
relationship is close to linear and comparable to a conventional fixed cavity process.
The triaxial and plain weave show a more uniform thickness distribution, regardless
of the time the inlet has been closed. In addition the thickness of the preforms does
not vary significantly with bleeding time. It highlights the lower compressibility -
relaxation behaviour of both triaxial and plain weave as demonstrated by fibre
content variations in Table 2.6 (see Equation 2.10). Fibre content was measured
using the burn off method. The latter consists of burning 50 by 25 mm samples (45
samples were burnt from each laminate produced) of laminate at 600
o
C for 45 min.
The fibre content is then obtained by calculating the volume of residu obtained and
reporting it to the volume of the sample, measured in water using Archimede law.
The volume of residu is obtained by reporting the mass of residu to the density of
the residu material (glass fibre in the present case).
- 41 -
Fig. 2.16. Thickness distribution in the different laminates.
Due to its high compressibility - relaxation behaviour CFRM shows the largest thickness
changes but the preform still does not show sign of a homogeneous thickness distribution after
20 min of bleeding. Plain weave presents unsignificant thickness changes while the triaxial
fabric shows low variations. These results show that the bleeding technique is not a successful
method to gain homogeneous thickness distribution in a preform and it only affects materials of
high compressibility - relaxation behaviour.
This study shows that bleeding is not necessarily a recommended method to end an
infusion. First it is only interesting for preforms of high compressibility - relaxation
level and does not tremendously affect other preforms of lower compressibility -
relaxation level.
Average thickness on
normalised width
1
2
3
4
1
2
3 4
3.5
4
4.5
5
5.5
6
6.5
7
7.5
0 1 2 3 4 5 6
Normalised distance from inlet
T
h
i
c
k
n
e
s
s
(
m
m
)
CFRM 20 min
CFRM 10 min
CFRM 2 min
FGE117 10 min
FGE117 2 min
FGE117 0 min
PW600 10 min
PW600 2 min
PW600 0 min
- 42 -
Results show that after 20 min the fibre content of a preform of high compressibility
- relaxation level has increased by 18 % and a uniform thickness distribution has
still not been reached. This shows that, although vacuum level is high, 101.3 mPa of
absolute pressure in the present case, a long time is necessary to obtain a uniform
thickness distribution and therefore the laminate is prone to resin starvation.
Fig. 2.17. Evolution of upper flow front position with time.
As expected infusion time decreases as the driving pressure gradient increases.
Table 2.6. Thickness and fibre content of each laminate
Reinforcement
Min thickness
(mm)
Max
thickness
(mm)
Min V
f
(%)
Max V
f
(%)
V
f
variation
(%)
U 751 / 375 20 min 4.249 5.563 23.9 31.3 23.6
U 751 / 375 10 min 4.746 6.592 20.2 28.0 28.0
U 751 / 375 2 min 5.210 6.769 19.6 25.5 23.0
FGE 117 10 min 6.888 7.408 55.9 60.1 07.0
FGE 117 2 min 7.073 7.843 52.8 58.6 09.8
FGE 117 0 min 6.869 7.434 55.7 60.3 07.6
RT 600 10 min 3.606 4.354 48.8 59.0 17.8
RT 600 2 min 3.489 4.162 51.1 60.9 16.6
RT 600 0 min 3.465 4.364 48.7 61.4 20.6
2.6. Conclusions
This general study of vacuum infusion shows that the flow enhancement properties
of the textured bag are due to a low conformation of the bag under vacuum more
than a real efficiency of the cracked ice pattern of the bag itself. Consequently the
0
50
100
150
200
250
300
350
400
450
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
e
c
)
101.3 mbar
401.3 mbar
701.3 mbar
T
i
m
e
(
s
)
- 43 -
use of a distribution medium with a smooth bag appears to be a preferable
combination.
The use of a breather felt is not recommended when combined with a distribution
medium as it masks the flow enhancement properties of the material by compacting
the preform futher.
As expected the use of a distribution medium in VI results in trans-plane flow in the
preform and infusion time is a function of permeability - porosity relationship (both
in-plane and trans-plane) as well as preform thickness. No dominant factor emerges
at this stage of the study but it is obvious that the size of lead-lag is directly linked
to the difference in permeability between the preform and the distribution medium
as well as the thickness of the preform.
Finally the control of cavity pressure and resin supply at the termination of the
infusion phase is seen to be critical in maintaining an homogeneous fibre content in
the parts whilst preventing resin starvation.
According to these preliminary results two key studies can be envisaged:
- A further series of statistically designed experiments focusing on the study of the
lead-lag followed by an experimental and computational analysis of the
distribution medium responsible for lead-lag occurence.
- The investigation on an integrated vacuum bag that meets the flow
enhancement properties of the distribution medium.
- 44 -
References
1. M. Feiler, L. Ischtschuk. Vacuum Assisted Resin Infusion (VARI): on the way to
serial production. 24
th
International SAMPE Europe Conference, April 2003, Paris,
France, p. 683-691.
2. M. Andersson, S. Lundstrom, B.R. Gebart, R. Langstrom. Development of
guidelines for the vacuum infusion process. 8
th
Fibre Reinforced Composite
Conference (FRC), Sep 13-15, 2000, Newcastle, UK, p. 113-120.
3. N.C Correia, F. Robitaille, A.C Long, C. D Rudd, P. Simacek, S.G Advani. Use of
resin transfer moulding simulation to predict flow, saturation and compaction in
the VARTM process. Journal of Fluids Engineering. ASME, 126 (2), 2004, p. 210-
215.
4. F. Robitaille, R. Gauvin. Compaction of textile reinforcement for composites
manufacturing. I: review of experimental results. Polymer Composites, 19, N
o
2,
April 1998.
5. A. Hammami, B.R. Gebart. A model for the vacuum infusion moulding process.
7
th
Fibre Reinforced Composite Conference (FRC), 1998, Newcastle, UK, p. 136-
145.
6. A. Hammami, B.R. Gebart. Model for vacuum infusion molding process. Plastics,
Rubber and Composites Processing and Application, 1998, 17, N4.
7. A. Hammami, B.R. Gebart. Experimental investigation of the vacuum infusion
moulding process. 8
th
European Conference on Composite Materials (ECCM-8),
June 1998, Naples, Italy, p. 577-584.
8. N. C. Correia. Analysis of the vacuum infusion moulding process. PhD thesis,
2004, University of Nottingham.
9. C. Williams, J. Summerscales, S. Grove. Resin Infusion under a Flexible Tooling
(RIFT): a Review. Composites Part A, 27, 1996. p. 517-524.
- 45 -
10. M. Thibaudeau, R.A. Shenoi. Understanding flow behaviour in the resin infusion
process. 10
th
European Conference on Composite Materials (ECCM-10), June 3-
7, 2002, Brugge, Belgium.
11. S.G. Advani et all. Flow and rheology in polymer composites manufacturing.
Newark, Elsevier, ISBN 0-444-89347-4 (Vol. 10), 1994. Chapter 12, p 465-515.
12. B. Qi, J. Raju, T. Kruckenberg, R. Stanning. A resin film infusion process for
manufacture of advanced composite structures. Composite Structures, 47, 1999,
p. 471-476.
13. A. Hoebergen, J. Holmberg. Vacuum infusion, ASM Handbook, 21, Composites,
Materials Park (OH), 2001, p. 501-515.
- 46 -
Chapter 3
Analysis of lead-lag in vacuum infusion
3.1. Introduction
The use of a high permeability layer to enhance flow in VI completely modifies the
impregnation kinetics by transforming an originally in-plane flow, when no medium
is used, into a flow principally in-plane or trans-plane at different locations. Results
from Chapter 2 show that differences in in-plane permeability between the
distribution medium and the preform introduce a delay, referred to as lead-lag,
between flow arrival on the upper and lower faces of the preform. The flow front in
the distribution medium progresses ahead of the flow front in the preform. Lead-lag
describes the state of flow front advancement on the upper and lower sides of the
preform as shown in Figure 3.1. It can be expressed as a time or a distance. Lead-
lag time, t
ll
(ll stands for lead-lag), characterises the time difference between flow
front arrival on the lower side and the upper side of the preform to reach a position
x (Equation 3.1). Lead-lag distance, x
ll
, represents the distance between flow fronts
on the upper and lower sides of the preform at any time t (see Equations 3.2 and
3.3). Both may vary as the infusion progresses (x
ll1
= x
ll1
).
1 2
t t t
ll
= 3.1
1 2
1
x x x
ll
= 3.2
2 2
1
x x x
ll
=
'
'
3.3
This chapter investigates the creation of lead-lag in vacuum infusion. The study
focusses on the effect of preform thickness (reinforcement layers, distribution
medium layers, preform compaction) and permeability on the occurrence of lead-
lag.
- 47 -
A way to control lead-lag during infusion is also investigated. It involves the cutting
the distribution medium shorter than the preform and analysing the progression of
lead-lag with flow front distance. Finally the work is completed by a series of
experiments designed for specific cases where the number of plies vary in a single
preform.
Fig. 3.1. Characterisation of lead-lag as a difference in time t or distance x.
3.2. Background
Lead-lag in VI results from the use of a flow enhancement layer on top of the
preform. The objective is to provide a high permeability layer to the low
permeability preform in order to enhance resin flow. The use of such material is not
x
1
t
1
x
2
Distribution medium
Outlet Inlet
P
1 P
2
Vacuum bag
Lower mould
Flow front
x
2
x
2
t
2
Distribution medium
Outlet Inlet
P
1 P
2
Vacuum bag
Lower mould
Flow front
Reinforcement
Reinforcement
- 48 -
new and flow dynamics in the presence of lead-lag have already been studied
experimentally and modelled.
As described by Seemann [1], a distribution medium is made of an array of spaced
lines, which can criss-cross with each other. It can also be a knitted or woven fabric
but in any case the distribution medium is preferably made of polymer
monofilaments which are non resin absorptive. Other authors have reported the use
of glass chopped strand mats or continuous filament random mats as distribution
medium layer [2, 3].
Several authors already studied lead-lag in VI. Thibaudeau [4] recorded flow
progression through the thickness of preforms made of unidirectional, woven or
randomly oriented textiles made of E-glass fibres. The type and permeability of the
distribution medium was not mentioned. The author pointed out the creation of a
lead-lag and the occurrence of flow through the thickness of the preform, also
mentioned by Searle [5] and Feiler [6]. Regrettably the length of the lead-lag was
not specified and no details about flow dynamics nor explanation on the creation of
lead-lag were given. Andersson [7] measured visually lead-lag during VI by colour
marking the different layers of the preform. The distribution medium was a chopped
strand mat of 300 g.m
-2
, and the preform a non-crimp stitch bonded reinforcement
of 800 g.m
-2
. It was found that lead-lag distance reached a maximum value which
occured near the inlet, and remained constant during most of the infusion. This was
confirmed by other authors [5]. But no explanation was offered to understand how
lead-lag is created and why it reached a maximum prior to a steady-state. Other
authors [8] showed that the flow front in the distribution medium lead the flow front
on the lower surface and therefore concluded that using a distribution medium
created a trans-plane flow in the preform and decreased infusion time. Sun [9]
confirmed this and added that infusion time was highly dependent on the
permeability of the distribution medium while it was not very sensitive to
permeability variations in the reinforcement.
Hammami [10] proposed to use a flow factor, which was function of the
permeability of the distribution medium related to that of the preform, in order to
determine the importance of trans-plane flow in the preform. The distribution
medium was assumed to behave as an open channel and trans-plane flow in the
preform was believed to progress vertically. However the latter was not confirmed
- 49 -
by validation. The flow model developed by Song [8], describing flow through a
distribution medium and a preform, assumed that the distribution medium could be
modelled as heterogeneous and anisotropic porous media. The model predicts
thickness changes in the preform as well as fill time. Results show to be in
agreement with experiments. Like Song, Han [11] also takes into account the
compressibility of the preform in his model for flow in SCRIMP which predicts the
occurrence of trans-plane flow.
Advani [12] modelled flow behaviour in RTM when infusing fabrics of different
permeability. His model, reported in Equation 3.4, considers that the permeability of
the stack is an average of the in-plane permeability of each fabric. The number and
thickness of each fabric layer is taken into account but trans-plane flow is
considered small compared to in-plane flow and only in-plane permeability is
involved. The average permeability value is then incorporated into Darcys law.

=
=
n
l
l
ij
l
ij
K h
H
K
1
1
3.4
This equation may be used to describe in-plane flow in VI when using fabrics of
different permeability. The average in-plane permeability is a direct function of the
number of fabric layers. Consequently, a constant ratio between the number of
distribution medium and reinforcement layers number should result in constant
infusion times (see chapter 2) [13].
However, trans-plane flow occurs in VI and trans-plane permeability of the preform
has to be considered in the model. The literature shows that trans-plane flow does
occur at the flow front. Behind the front flow is essentially in-plane. Advanis
updated model for flow in RTM [12] proposes that behind the front no trans-plane
flow occurs and the average in-plane permeability model can be used. At the front,
trans-plane flow does occur. Consequently it is assumed that the flow splits in two
regions, one where the average in-plane permeability model is used and one where
a flow front permeability K
ff
is used, as described in Equation 3.5 and shown in
Figure 3.2. In the latter case the permeability is a function of both individual fabric
in-plane permeability and fabrics average in-plane permeability. The model does not
depend directly on trans-plane permeability, but the length of the flow front region
where this model is valid does depend on it. Advani assumes that the pressure drop
- 50 -
at this flow front region is a function of trans-plane permeability since it depends on
the ratio K
ff
/Al as described by Darcys law.
a
f
f
ff
K
K
K
K
|
|
.
|

\
|
+ = 1
2
1
2
1
3.5
K
ff
is the average permeability at the front, K
ij
the average permeability behind the
flow front and K
f1,2
the permeability of the differentmaterials infused.
Fig. 3.2. Sections of the preform where the average permeability model
and transverse flow model can be used, according to Advani [11].
This model gives a more realistic expression of flow in RTM when fabrics of different
inplane permeability are impregnated. However the use of such a model to
describe flow in VI has to be demonstrated. Inplane and trans-plane permeability
variations of the wet preform with flow front progression have to be measured. An
updated version of the model [14] suggests a parabolic flow front and holds that
lead-lag is constant and can be calculated. Flow predictions were not compared to
experiments and the model was not validated.
Previous models and experimental analysis of flow in VI show that trans-plane flow
occurs in the preform when a flow enhancement medium is used (see chapter 2).
Numerous assumptions have been made in models but the physics of the
impregnation has not been analysed in detail. The process for lead-lag creation and
its evolution during the impregnation of the preform is not fully understood. Two
Outlet
P
1
Inlet
P
2
Upper mould
Lower mould
K
ij
K
ff
Distribution medium
Dry area
Flow front
Wet area
Reinforcement
- 51 -
aspects of the problem can be studied experimentally in order to better understand
lead-lag:
- First determine the importance of the preform permeability, permeability
variations and thickness on lead-lag creation.
- Second measure the effect of the distribution medium permeability and
thickness on lead-lag creation and determine whether the distribution medium
behaves as an open channel or a porous medium.
The control of lead-lag during VI was also investigated and consequences on
preform impregnation and infusion time were made. Finally flow through preforms
of less conventionnal lay-up, i.e increase or decrease of plies number in a single
preform as well as overlapping, was studied experimentally and the effect on lead-
lag was analysed.
In order to achieve these objectives, three types of experiment were conducted:
one varied the number of reinforcement layers while keeping a constant number of
distribution medium plies. The second investigated the possibility of eliminating
lead-lag in VI by imposing a cut-off distance to the distribution medium, on the
outlet side. The third varied the number of reinforcement layers in a single preform
and involved overlaps in preforms of constant number of layers.
3.3. Assumptions
3.3.1. Flow front recordings
During the experiments, the times at which the flow front reached points located 50
mm apart on the centre line of the preform were recorded for both the upper and
lower faces of the preform (0 mm = position of the inlet). Regarding the lower face
of the preform, a distinction must be made between the recorded positions of the
front and the actual flow rate of the fluid particles during the infusion. Figure 3.3
highlights the progression of the flow front with time between 425 mm and 490 mm
for a 9-layer preform with 0 tow oriented perpendicular to the flow direction. In this
example it takes 172 seconds for the flow front to progress through the 65 mm
separating points A and C on the lower side of the preform. The curve shows that at
the same time the front progresses from point B to point C through the thickness.
If it is assumed that trans-plane flow leads at the front and in-plane flow leads
- 52 -
behind the front, it is unlikely that velocity vectors are changing immediately from
horizontal orientation in the in-plane flow to vertical orientation in the trans-plane
flow.
Fig. 3.3. The difference between recorded flow and actual flow in VI.
B
Outlet
Inlet
x
z
Flow front
C
A
x
A C
B
Distance recorded
z
0
200
400
600
800
1000
1200
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
e
c
)
Test 3 - Upper
Test 3 - Lower
B
C
A
425 mm 490 mm
587 s
759 s
)
Lower flow front
position at 759 s
Lower flow front
position at 587 s
Upper flow front
position at 587 s
- 53 -
There must be an evolution of velocity vector orientation from flow behind the front
to flow at the front. Therefore it is not obvious whether at the front velocity vectors
are exactly perpendicular to the plane of the preform or not. If it is assumed,
following Advani [13], that the flow front is parabolic many solutions exist to
represent velocity vector orientations. Therefore times recorded on the lower face of
the preform correspond to times for an apparent in-plane distance covered which is
not the real distance covered by the fluid.
Lead-lag expressed as a difference in time represents the time required for the flow
on the lower side of the preform to impregnate a distance already reached by the
flow on the upper side of the preform (see Equation 3.1). During VI experiments,
times at which the front reached points located 50 mm along the centreline of the
preform were recorded for both the upper and lower faces of the preform (0 mm =
position of the inlet). According to the above explanation it is not possible to
express lead-lag as a difference in time since times recorded on the lower face of
the preform were recorded for successive covered distances of 50 mm, different
from the real distance covered by the fluid between two successive measuring
points.
Still, it has to be demonstrated whether trans-plane flow is exactly perpendicular to
the plane of the preform or not. One solution consists of comparing the theoretical
and experimental thickness of the preform. The theoretical thickness is calculated
using data of trans-plane permeability, compaction and porosity of the preform.
Trans-plane permeability measurements give permeability data of the fabric for a
range of fibre contents. Assuming that the pressure distribution at the flow front is
known, the corresponding trans-plane permeability and fibre content distribution is
obtained using compaction data of the fabric. Therefore a thickness distribution of
the preform can be determined. The real distance x covered by the fluid between
two recorded flow front positions is determined using an integration of Darcys law,
described in Equation 3.6.
P
x
K
t
A

=
2
2
c u
3.6
Assuming that the porosity and trans-plane permeability of the preform are known
- 54 -
for the corresponding pressure distribution then x can be determined using times
recorded for the different positions of the flow front. Theoretical thickness and flow
distance x are then compared. If results are similar then it is proved that trans-
plane velocity vectors are perpendicular to the preform. In the other case velocity
vectors are not perpendicular to the preform and the distance covered by the fluid is
unknown. A direct consequence is that the velocity expressed at the lower side of
the preform is apparent and not real.
3.3.2. Creation of lead-lag
The occurrence and evolution of lead-lag in VI is assumed to depend on two main
factors:
- The permeability of both the distribution medium and the reinforcement
- The thickness of the preform
Both in-plane and trans-plane permeability of the preform should influence lead-lag
since, according to Advani, in-plane flow occurs behind the front as well as trans-
plane flow at the front.
The thickness of the preform is assumed to influence the length of the lead-lag.
Indeed the distance to flow through the preform increases with stack thickness.
Therefore lead-lag should increase with preform thickness.
Consequently lead-lag should be a function of preform thickness, in-plane
permeability of the preform and both in-plane and trans-plane permeability of the
distribution medium.
However the influence of permeability variations in the reinforcement on lead-lag is
not known and Suns statement, which specifies that infusion time is highly
dependent on the permeability of the distribution medium but not very sensitive to
permeability variations in the reinforcement, has to be verified [9].
3.3.3. Control of lead-lag
The presence of a lead-lag in VI is not an issue as long as the preform is completely
impregnated before gelation occurs. However there are some cases where lead-lag
is not desirable. For example, infusion strategies which involve sequential opening
- 55 -
and closing of inlets require a lead-lag free flow in order to avoid dry patches in
the preform (see Figure 3.4). When the flow reaches a new inlet, the pressure
gradient is applied between the new inlet and the outlet.
a). Use of a distribution medium. Inlet 1 is closed when upper flow reaches Inlet 2.
Inlet 2 is opened: a dry patch occurs in the preform.
b). No distribution medium is used. Both upper and lower flow reach
Inlet 2 at the same time. Inlet 2 is opened and Inlet 1 is closed.
Fig. 3.4. Effect of lead-lag on preform impregnation
when sequential opening strategy is used.
No pressure gradient exists between the new inlet and the previous one so no
pressure difference drives the resin in that section of the preform.
In theory if a dry patch occurs it cannot be impregnated at a later stage. However it
is assumed that the dry patch is slightly impregnated by capillarity but, because it is
a slow process, results in high resin consumption. An example of this situation is
shown in Figure 3.4 a).
One solution is to eliminate lead-lag before the upper flow front reaches the next
inlet. Because lead-lag is due to a difference in permeability between the
distribution medium and the reinforcement the method consists of cutting the
Inlet 1 Inlet 2 Outlet
No lead-lag
No dry patch
Inlet 1 Inlet 2 Outlet
Lead-lag
Dry patch
- 56 -
distribution medium shorter than the preform so that only reinforcement
permeability governs flow progression. Consequently in-plane flow is expected to
progressively replace trans-plane flow and lead-lag to disappear since no
permeability difference exists in the preform. The best distance for the cutting of
the distribution medium has to be determined as a function of the time required for
lead-lag to disappear.
3.4. Experimental analysis: materials and method
3.4.1. Materials
Technical data of materials and fluid are available in Appendix 2.
Reinforcement
A triaxial E-glass stitched fabric E-TLX 1169 (0 / 45, 1169 g/m
2
) from Cotech
was used for the experiments.
Fluid
The fluid consisted of pre-accelerated isophthalic polyester resin 701 PAX from Scott
Bader. It was initiated with 1 wt% of Butanox M50 from Akzo Chemie.
Consumables
The materials also included a smooth polyamide vacuum bag (B205, 0.05 mm thick,
Tygavac) and a polyamide woven bleed out fabric (60B from Tygavac) used to
release other bagging materials from the cured part. The distribution medium
consisted of a high density polyethylene (HDPE) warp knitted chained stitch textile
with inserted filaments (H8036, 100 g/m
2
, filament diameter = 0.25 mm, Palmhive
Textiles). The architecture of the distribution medium is shown in Figure 3.5.
3.4.2. Method
The materials were laid on a 1 m by 1 m glass plate and the components produced
measured 0.2 m by 1.0 m as they were set along the diagonal of the glass plate.
The moulding set-up is shown in Figure 3.6 and 3.7. The air was evacuated from the
cavity using an Edwards pump, model RV5. The level of vacuum was adjusted on
the pump to 101.3 mbar of absolute pressure. It was checked on a vacuum gauge
- 57 -
connected to the pump through a digital control screen. Steel springs (15 mm
diameter) were used along the inlet and outlet edges to promote rectilinear flow
throughout the cavity.
Fig. 3.5. Architecture of the distribution medium
Fig. 3.6. Materials assembly
3.5. Experimental analysis: results
3.5.1. Creation of lead-lag in VI
The experiments involved the moulding 0.2 m by 1.0 m flat plaques made of E-TLX
1169 and varying the number of reinforcement layers while keeping one layer of
distribution medium on top, as described in Table 3.2. The distribution medium was
oriented warpwise to the direction of the flow. Each test was conducted three times
to assess the repeatability of the process.
Vacuum bag
Distribution medium
Sealant tape
Mould
Steel sping
Outlet Inlet
Reinforcement
Bleed out fabric
Warp (inserted filament)
Weft (stitch)
- 58 -
Fig. 3.7. VI experimental set-up
Table 3.2. Test matrix: effect of preform thickness on lead-lag
Reinforcement Number of plies
Distribution medium
layers
0 tow / flow
direction
3 1 Parallel
6 1 Parallel E-TLX 1169
9 1 Parallel
Figure 3.8 shows the evolution of flow front distance squared as a function of
infusion time for 3, 6 and 9 triaxial layers preforms. The orientation of the 0
o
tow is
parallel to the direction of flow. Results for flow on both upper and lower faces of
the preform are presented. The lines are polynomial curve fits.
Results show that the difference in infusion time between the upper and lower side
of the preform increases with preform thickness. Figure 3.9 shows lead-lag distance
x
ll
as a function of flow front position on the upper side of the preform. x
ll
increases
with flow front progression and tends to a steady-state. Since no lead-lag exists
before infusion starts, x
ll
must go through a maximum value. In addition x
ll
increases with preform thickness.
Resin trap
Pump
Preform and bagging materials
Spring (outlet)
Glass plate
Steel spring (inlet)
- 59 -
Fig.3.8. Flow front position squared as a function of time.
Trendlines are polynomial curve fit.
As expected infusion time increases with preform thickness. In addition the delay between upper
and lower flow increases with preform thickness.
Fig.3.9. Lead-lag expressed as a difference in distance.
Trendlines are polynomial curve fit.
Lead-lag increases with preform thickness due to a greater trans-plane distance to flow through.
Generally lead-lag reaches a maximum near the inlet and decreases to a steady-state.
The evolution of lead-lag is confirmed by the velocity of the fluid on both upper and
lower side of the preform as shown in Figure 3.10. At the onset of the infusion, the
velocity of the flow front on the upper wall is significantly higher than on the lower
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
8.0
9.0
0 500 1000 1500 2000 2500 3000
Time (sec)
F
l
o
w
f
r
o
n
t
p
o
s
i
t
i
o
n
s
q
u
a
r
e
d
(
1
0
5
m
m
2
)
3 layers upper
3 layers lower
6 layers upper
6 layers lower
9 layers upper
9 layers lower
)
0
10
20
30
40
50
60
70
80
90
0 200 400 600 800 1000
Flow front position upper (mm)
L
e
a
d
-
l
a
g
(
m
m
)
3 layers
6 layers
9 layers
- 60 -
wall. However the velocity of the flow on the upper side decreases more quicker
than the velocity on the lower side until both become similar as shown in Figure
3.11.
Fig. 3.10. Fluid velocity as a function of flow front position
on both upper and lower faces of the preform.
Velocity is higher in the upper face of the preforms and decreases on both sides with preform
thickness.
The fact that both lead-lag and infusion time increase with plies number shows the
dependence of the impregnation dynamics on preform thickness. This is due to the
occurrence of a trans-plane flow. At the start of the infusion, the flow splits into in-
plane and trans-plane components. Figure 3.12 shows the flow front distance, on
both upper and lower sides of the preform, at which maximum lead-lag x
llmax
occurs
and as a function of the number of reinforcement layers.
Figure 3.13 shows the time at which x
llmax
occurs for an increasing number of
reinforcement layers. x
llmax
increases with preform thickness. In addition the flow
front distance and time at which x
llmax
occurs increases on both upper and lower
sides with preform thickness. These results are explained as follows: at the start of
the infusion in-plane flow occurs in both the distribution medium and the
reinforcement.
0
2
4
6
8
10
12
14
0 200 400 600 800 1000
Flow front position (mm)
V
e
l
o
c
i
t
y
(
m
m
/
s
e
c
)
3 layers upper
3 layers lower
6 layers upper
6 layers lower
9 layers upper
9 layers lower
V
e
l
o
c
i
t
y
(
m
m
/
s
)
- 61 -
Fig. 3.11. Fluid velocity as a function of flow front position on both upper and lower
faces of the preform between 490 mm and 910 mm from the inlet.
Trendlines are polynomial curve fit.
After the start of the infusion velocity decreases sharply and upper velocity tends to meet lower
velocity value and this whatever the thickness of the preform.
Fig.3.12. Flow front distance from inlet on both upper
and lower faces of the preform when maximum lead-lag occurs.
The upper and lower flow front distance at which x
llmax
occurs increases with preform thickness
due to the greater trans-plane distance to cover for thicker preforms.
0
50
100
150
200
250
0 3 6 9
Number of reinforcement layers
D
i
s
t
a
n
c
e
f
r
o
m
i
n
l
e
t
(
s
e
c
)
Upper flow front distance
Lower flow front distance
L
d
m
a
x
0.0
0.2
0.4
0.6
0.8
1.0
450 550 650 750 850 950
Flow front position (mm)
V
e
l
o
c
i
t
y
(
m
m
/
s
e
c
)
3 layers upper
3 layers lower
6 layers upper
6 layers lower
9 layers upper
9 layers lower
)
D
i
s
t
a
n
c
e
f
r
o
m
i
n
l
e
t
(
m
m
)
V
e
l
o
c
i
t
y
(
m
m
/
s
)
- 62 -
Fig.3.13. Time at which maximum lead-lag occurs
according to the number of reinforcement layers.
Because the in-plane permeability of the distribution medium is much higher than
that of the reinforcement, flow in the distribution medium (upper side of the
preform) is ahead of the flow in the reinforcement, as recorded on the lower side of
the preform.
That is how lead-lag is created. Trans-plane flow occurs at the start of the infusion.
Flow in the distribution medium splits immediately at the onset of the infusion into
in-plane and trans-plane flows, and fluid velocity vectors gradually change their
orientation from in-plane direction into trans-plane direction, as shown in Figure
3.14.
Fig. 3.14. Schematic of split of in-plane flow to trans-plane flow.
Evolution of velocity vector orientations.
3 layers
6 layers
9 layers
0
10
20
30
40
50
60
70
80
90
35 45 55 65 75 85 95
L
dmax
(mm)
T
i
m
e
(
s
e
c
)
Distribution medium
Preform
Velocity vector
T
i
m
e
(
s
)
x
llmax
- 63 -
However, until trans-plane flow reaches the lower side of the preform for the first
time, recordings of flow front progression with time on the lower face of the mould
correspond to in-plane flow in the reinforcement only, as shown by the blue section
in Figure 3.15. Figure 3.16 shows a Log / Log representation of lead-lag as a
function of time from the start of the infusion to the occurrence of x
llmax
.
Flow front progression in both the distribution medium and the reinforcement. In-
plane flow still dominates in the reinforcement; trans-plane flow has not yet reached the
lower side of the preform.
Time at which x
llmax
occurs. Trans-plane flow reaches the lower side of the preform for
the first time.
Fig. 3.15. Evolution of lead-lag with time and occurrence of x
llmax
.
Experimental results for a 3-layer preform, curve fit up to x
llmax
.
0
5
10
15
20
25
30
35
40
0 50 100 150 200 250 300 350 400
Time (sec)
L
e
a
d
-
l
a
g
(
m
m
)
2
5
m
m
25 mm
3
8
m
m
38 mm
Inlet Outlet
In-plane flow
Distribution medium
In-plane flow
Trans-plane flow
Reinforcement
Experimental data
Data fit
)
- 64 -
Fig. 3.16. Log (lead-lag) as a function of Log (Time) between
the start of the infusion to the occurrence of x
llmax
.
x
llmax
occurs when trans-plane flow reaches the lower side of the preform for the
first time, which corresponds to the red section in Figure 3.15. This explains why
the upper and lower flow front distances reached when x
llmax
occurs increase with
the number of reinforcement layers as it takes a longer time for the fluid to flow
through thicker preforms. This explanation is confirmed by the fact that, at the
beginning of the infusion, both upper and lower flow front positions with respect to
time do not follow a linear trend and flow front distances squared are not
proportional to time, as shown in Figure 3.16. This means that the permeability of
the preform evolves with both flow front progression and pressure gradient.
Otherwise, as stated by Equation 3.6, time would be proportional to flow front
distance squared. However after the start of the infusion, the lower flow front
distance squared is proportional to time but not the upper flow front distance and
the non-linear trend increases with preform thickness (see Figure 3.8). This means
that trans-plane permeability of the reinforcement is not influenced by the pressure
gradient and remains constant (see Appendix 5). The in-plane permeability does not
lead the impregnation of the reinforcement as it did before. The distribution medium
is impregnated by in-plane flow as the curve demonstrates the influence of in-plane
permeability changes with pressure gradient by showing a non-linear trend.
1.0
1.1
1.2
1.3
1.4
1.5
1.6
1.7
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Log (Time (s))
L
o
g
(
L
e
a
d
-
l
a
g
(
m
m
)
)
- 65 -
Fig. 3.16. Flow front position squared as a function of time, at the start of the
infusion. Trendlines are polynomial curve fit.
Upper and lower flow front distances squared are not proportional to time. Therefore preform
permeability evolves with both flow front progression and pressure gradient.
The fact that in-plane permeability leads during the first phase of the impregnation,
i.e until maximum lead-lag occurs, is confirmed by the following test. It consists of
comparing the impregnation of two preforms where the orientation of the 0 tow to
flow direction was turned from parallel in the first preform to perpendicular in the
second preform (see Table 3.3). Figure 3.17 and 3.18 show the results for 3 and 9-
layer cases. The distribution medium was oriented weftwise to the direction of the
flow.
Table 3.3. Test matrix: effect of material orientation
Reinforcement Number of plies 0 tow / flow direction
3 Parallel
9 Parallel
3 Perpendicular
E-TLX 1169
9 Perpendicular
Results show that, after 500 mm of flow progression, a difference of 67 sec is
recorded on the top of the 3-layer preforms, the quicker flow being in the preform
with the 0 tow oriented parallel to the flow direction. On the bottom of the
preforms, the difference is 42 sec. In the 9-layer preform with the 0 tow oriented
parallel to the flow direction, flow front reaches 500 mm on the upper side of the
preform 168 sec ahead of the other 9-layer preform.
0.0
0.1
0.2
0.3
0.4
0.5
0 20 40 60 80 100 120 140 160 180
Time (sec)
F
l
o
w
f
r
o
n
t
p
o
s
i
t
i
o
n
s
q
u
a
r
e
d
(
1
0
5
m
m
2
)
3 layer upper
3 layers lower
6 layers upper
6 layers lower
9 layers upper
9 layers lower
)
- 66 -
Fig. 3.17. Infusion time versus flow front position. 0 tow parallel to flow direction.
Trendlines are polynomial curve fit.
Fig. 3.18. Infusion time versus flow front position. 0 tow perpendicular to flow direction.
Trendlines are polynomial curve fit.
The bottom of the preforms shows a difference of 198 sec. In addition in-plane
permeability measurement of a 44 % fibre volume fraction triaxial fabric shows a
transient permeability for flow parallel to the 0 tow of 1.06*10
-11
m
2
against
7.65*10
-12
m
2
in the perpendicular direction (see Appendix 1). This demonstrates
the strong effect of fabric orientation on in-plane flow and the great importance of
0
200
400
600
800
1000
1200
1400
1600
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
e
c
)
3 layers upper
3 layers lower
9 layers upper
9 layers lower
T
i
m
e
(
s
)
0
200
400
600
800
1000
1200
1400
1600
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
e
c
)
3 layers upper
3 layers lower
9 layers upper
9 layers lower
)
T
i
m
e
(
s
)
- 67 -
in-plane permeability in VI even if trans-plane flow leads at the front. It confirms
the lead of in-plane flow in the preform during a phase of the impregnation and its
influence on lead-lag distance.
Figure 3.19 shows the lead-lag obtained for the different tests. As expected lead-lag
is greater for the 9-layer preforms. The effect of fabric orientation, and thus in-
plane permeability, on lead-lag is demonstrated as lead-lag is greater when the 0
o
tow is oriented parallel to the flow direction. This result is expected as it confirms
what obtained in in-plane permeability measurements, i.e the difference of flow
front advancement is a function of the orientation of the fabric to the flow direction.
Fig. 3.19. Lead-lag expressed as a difference in distance.
Trendlines are polynomial curve fit.
As expected lead-lag increases with plies number. In addition lead-lag is greater when the 0
o
tow
is oriented parallel to the flow due to a higher permeability of the fabric in that direction.
3.5.2. Control of lead-lag
Experiments were conducted to investigate the elimination of lead-lag by cutting the
distribution medium shorter than the preform. The idea was that, since the
distribution medium layer was cut, fluid would flow through the reinforcement only.
Therefore trans-plane flow might disappear and in-plane flow recover the lead to
impregnate the preform. The time and distance necessary to recover in-plane flow
were analysed as a function of preform thickness related to distribution medium
thickness. The experiments were performed by testing 3 different lay-ups as
described in Table 3.4.
0
20
40
60
80
100
120
140
0 100 200 300 400 500 600
Flow front position lower (mm)
L
e
a
d
-
l
a
g
(
m
m
)
3 layers perpendicular
3 layers parallel
9 layers perpendicular
9 layers parallel
- 68 -
Table 3.4. Test matrix: elimination of lead-lag in VI
Reinforcement Distribution medium
Plies
number
0 tow/flow
direction
Length (mm)
Plies
number
Orientation Length (mm)
8
11 Parallel 700 1 Warpwise 500
14
The triaxial preform was 700 mm long and the distribution medium was cut 200 mm
shorter. Recordings of infusion times with flow front progression were made every
50 mm, 0 mm being at the inlet. After the cut-off occurred, time was recorded
every 10 mm.
Results in Figure 3.20 and 3.21 show that the time necessary to recover in-plane
flow increases with preform thickness and is 204 sec for the 8 layer preform, 244
sec for the 11 layer preform and 293 sec for the 14 layer preform.
Fig. 3.20. Time as a function of flow front position
on both upper and lower faces of the preform.
The time required to recover in-plane flow in the preform increases with preform thickness.
Before the cut in the distribution medium occurs, lead-lag is longer for thicker
preforms, as expected and shown in Figure 3.22. After the flow front reaches the
cut-off, lead-lag decreases until it completely disappears.
Figure 3.23 shows the decrease and elimination of lead-lag for the three preforms.
0
200
400
600
800
1000
1200
1400
1600
1800
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
e
c
)
8 layers upper
8 layers lower
11 layers upper
11 layers lower
14 layers upper
14 layers lower
T
i
m
e
(
s
)
- 69 -
Fig. 3.21. Time as a function of flow front position on both upper
and lower faces of the preform after cut-off occured.
The flow front distance at which in-plane flow recovers does not vary with preform thickness.
Fig. 3.22. Lead-lag expressed as a difference in distance until 450 mm of flow front
progression on the lower side of the preform, before cut-off occurs.
Trendlines are polynomial curve fit.
As expected lead-lag decreases after cut-off of the distribution medium occurred. Over a similar
flow front progression distance, lead-lag decreases of a greater amount with thicker preforms.
Results show that the distance necessary to eliminate lead-lag is similar for the
40
45
50
55
60
65
70
75
80
100 150 200 250 300 350 400 450 500
Flow front position (mm)
L
e
a
d
-
l
a
g
(
m
m
)
8 layers
11 layers
14 layers
800
900
1000
1100
1200
1300
1400
1500
1600
1700
500 510 520 530 540 550 560 570
Flow front position (mm)
T
i
m
e
(
s
e
c
)
8 layers upper
8 layers lower
11 layers upper
11 layers lower
14 layers upper
14 layers lower
)
T
i
m
e
(
s
)
- 70 -
three preforms and therefore is practically independent of the thickness of the
preform. This result can be explained as follows: the pressure difference over the
lead-lag distance is greater for thicker preforms as shown in Figure 3.24. Let us
assume that each preform has a given thickness h
1
, h
2
and h
3
with h
1
< h
2
< h
3
.
Cut-off occurs at time t
1
and to h
1
corresponds a lead-lag x
ll1
, to h
2
a lead-lag x
ll2
and to h
3
a lead-lag x
ll3
with x
ll1
< x
ll2
< x
ll3
. In order to eliminate lead-lag a greater
in-plane distance has to be covered in thicker preforms.
The pressure gradient between the flow front at the lower side of the preform and
the flow at the same flow front distance on the upper side of the preform is higher
for thicker preforms as shown in Figure 3.24.
Fig.3.23. Lead-lag expressed as a difference in distance after cut-off occured.
Trendlines are polynomial curve fit.
If the same distance is necessary to eliminate lead-lag whatever the thickness of
the preform, it means that the pressure gradient AP related to the length of the
lead-lag x
ll
is identical for all preforms.
3.5.3. Specific cases
The lay-up phase, prior to infusion, is not a straightforward exercise. Depending on
the geometry of the mould, reinforcement plies might overlap. In addition,
0
2
4
6
8
10
12
14
16
18
20
490 500 510 520 530 540 550 560
Flow front position (mm)
L
e
a
d
-
l
a
g
(
m
m
)
8 layers
11 layers
14 layers
- 71 -
depending on the mechanical properties required, the laminates can present varying
number of reinforcement plies in their lay-up. These changes can disturb flow and
modify lead-lag during the infusion. A series of experiments was conducted to
investigate their effect on preform impregnation and lead-lag evolution.
Fig. 3.24. Representation of the flow front position when cut-off occurs
in the distribution medium and for different preform thickness.
x
2
P
1
= P
inj
Inlet
Inlet
Inlet
x
1
t
1
x
2
Outlet
P
2
= 0
Vacuum bag
Lower mould
Flow front
Reinforcement
t
1
Outlet
P
2
= 0
x
3
t
1
x
2
Outlet
P
2
= 0
x
4
P
1
= P
inj
P
1
= P
inj
x
x
P
AP
1
x
2 x
1
AX
1
P
AP
2
x
2 x
3
AX
2
x
P
AP
3
x
2
x
4
AX
3
- 72 -
Varying reinforcement layers in a single moulding
Experiments were conducted where the number of reinforcement layers varied in a
single preform. The number of plies presented either an increase or a decrease
according to the direction of flow as shown in Figure 3.25. Table 3.5 describes the
tests conducted. Thickness changes were set every 50 mm, starting 300 mm from
the inlet.
Results were compared to infusion of preforms with constant number of plies.
Results are presented for 8-14 layers and 14-8 layers preforms. Infusion times for
the other preforms analysed are available in Appendix 6.
Figure 3.26 shows infusion time with flow front position for an 8 to 14 E-TLX 1169
plies increase. Results show that the difference in infusion time between the upper
and lower sides of the preform keeps increasing with flow front progression, and as
expected the upper flow front is ahead of the lower flow front.
a). Thickness increase b). Thickness decrease
Fig. 3.25. Effect of thickness changes on lead-lag. Lay-up details.
Table 3.5. Test matrix: varying number of reinforcement layers in the preform
Number of plies 0 tow / flow direction
8 to 11 Parallel
8 to 14 Parallel
11 to 14 Parallel
11 to 8 Parallel
14 to 11 Parallel
14 to 8 Parallel
8 Parallel
11 Parallel
14 Parallel
Inlet
0 fibre
axis
Outlet
Inlet
Outlet
0 fibre
axis
- 73 -
Figure 3.27 confirms the results and shows that lead-lag distance x
ll
increases with
thicker preforms. The 8-11 and 11-14 layer preforms follow the same tendency due
to a similar increase in plies number. As expected the 11-14 layer preform presents
a larger lead-lag due to a greater thickness at the onset of the infusion. The 8-14
layer preform presents the same lead-lag as the 8-11 layer preform at the start of
the infusion. Nevertheless it increases more with flow front progression than the
other preforms due to a greater increase in plies number (6 plies against 3 plies).
Figure 3.26 compares the infusion time of the 8-14 layer preform with preforms of
constant plies number, respectively 8 layers and 14 layers. Results show that the
infusion time on the upper side of the 8-14 layer preform follows the tendency of
the infusion time on the upper side of the 8-layer preform.
Fig. 3.26. Time as a function of flow front position on both upper and lower faces of the
preform. Comparison between preforms of constant thickness and preforms of increasing
thickness. Trendlines are polynomial curve fit.
Similarly the infusion time on the lower side of the 8-14 layer preform tends to
follow the same infusion time as the lower side of the 14-layer preform. This result
is expected since the thickness of the preform increases with distance from the
inlet. The infusion time on the upper side of the preform shows that the flow rate in
the distribution medium is not affected by the increase in preform thickness. This is
in contradiction with previous results. Figures 3.28 and 3.29 show respectively
infusion time and leadlag for a preform of decreasing plies number.
0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 100 200 300 400 500 600 700
Flow front position (mm)
T
i
m
e
(
s
e
c
)
8/14 layers upper
8/14 layers lower
Polynomial (14 layers upper)
Polynomial (14 layers lower)
Polynomial (8 layers upper)
Polynomial (8 layers lower)
)
- 74 -
Fig.3.27. Lead-lag expressed as a difference in distance.
Trendlines are polynomial curve fit.
Lead-lag on 8-11 and 11-14 layer preform follow the same tendency due to a similar increase in
number of plies.
Fig. 3.28. Time as a function of flow front position on both upper and lower faces of the
preform. Effect of thickness increase on infusion time. Trendlines are polynomial curve fit.
11-8 layer, 14-11 layer and 14-8 layer preforms were tested. Figure 3.28 presents
the results for the 14-8 layer preform. The infusion times of the other preforms are
available in Appendix 6. Infusion times between upper and lower side of the preform
60
70
80
90
100
110
120
130
140
150
250 300 350 400 450 500 550 600
Flow front position (mm)
L
e
a
d
-
l
a
g
(
m
m
)
8/11 layers
8/14 layers
11/14 layers
0
200
400
600
800
1000
1200
1400
1600
1800
2000
0 100 200 300 400 500 600 700
Flow front position (mm)
T
i
m
e
(
s
e
c
)
14/8 layers upper
14/8 layers lower
Polynomial (14 layers upper)
Polynomial (14 layers lower)
Polynomial (8 layers upper)
Polynomial (8 layers lower)
)
- 75 -
tend to a constant difference in time which is similar to what was obtained with
preforms of constant number of layers.
Fig. 3.29. Lead-lag expressed as a difference in distance.
Trendlines are polynomial curve fit.
Lead-lag decreases and as expected the same trend is observed for preforms with
similar decrease in plies number. Preforms of higher thickness present a higher
lead-lag. In addition the 14-8 layer preform presents a lower decrease of lead-lag
than the other preforms. This is explained as follow: let us compare this preform
with the 14-11 layer preform. When flow on the upper side of the 14-11 layer
preform reaches the 11 layers section of the lay-up, no other thickness change
occurs in the lay-up until the end of the infusion. Flow on the lower side of the
preform is still in between 14 and 11 layers, but because of the decrease in
thickness, trans-plane flow velocity increases and the lead-lag distance between
upper flow front and lower flow front decreases. It is expected that, at some point,
when both upper and lower flow are in the 11 layers section, lead-lag tends to a
steady state. In the present case though the preform was not long enough to record
this expected result before the end of the infusion.
In the 14-8 layer preform, the flow on the upper side of the preform reaches a
section with constant number of layers (8) only 100 mm before the end of the
preform length (the thickness changes start 300 mm after the beginning of the
infusion and every 50 mm a layer of reinforcement is removed) compared to 450
70
80
90
100
110
120
130
250 300 350 400 450 500 550 600
Flow front position (mm)
L
e
a
d
-
l
a
g
(
m
m
)
11 to 8 layers
14 to 11 layers
14 to 8 layers
- 76 -
mm in the 14-11 layer case and 11-8 layer case. This constant change in preform
thickness explains why lead-lag is constantly high.
This study shows that, in order to reduce infusion time and lead-lag when variations
in preform thickness occur, preforms have to be infused in their thickness decrease
direction.
Investigating overlapping of fabric layers in a preform
Overlapping reinforcement plies in a single preform was studied experimentally in
order to measure its effect on infusion time. 8-layer preforms with 50 mm overlaps
located every 50 mm (starting 50 mm from the inlet) were infused. Two cases were
studied: one consisted of placing an overlap on each layer and superimposing the
overlaps every 4 layers (see Figure 3.30). The other case featured overlaps on each
layers but no overlaps were superimposed (see Figure 3.31).
Fig. 3.30. Case A: repeatability of the overlaps: every 4 layers, no interval.
Fig 3.31. Case B: repeatability of the overlaps: every 4 layers, with 50 mm interval.
Inlet
Outlet
Reinforcement
Overlap
50mm 50mm
Inlet
Outlet
Overlap
50mm 50mm
Reinforcement
- 77 -
Table 3.6 describes the tests conducted. Results were compared to the infusion of
an 8-layer preform with no overlap.
Table 3.6. Test matrix: effect of overlaps on lead-lag
Reinforcement Overlap
Number of
plies
0 tow / flow
direction
Overlap / flow
direction
Superimposition
Interval every 4
layers
8 Perpendicular Perpendicular Every 4 layers
-
8 Perpendicular Perpendicular
-
50 mm
8 Perpendicular
- - -
Figure 3.32 shows infusion time as a function of flow front position for preforms
presenting an overlap and compares the results to infusion time for preforms with
no overlap.
Fig. 3.32. Time as a function of flow front position on both upper and lower face of
the preform. Effect of overlap on infusion time.
All trend lines are polynomial curve fit.
Overlaps result in longer infusion times due to local increases of plies number in the preform.
0
200
400
600
800
1000
1200
1400
1600
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
)
Upper no interval
Lower no interval
Upper interval
Lower interval
Polynomial (Upper no overlap)
Polynomial (Lower no overlap)
)
- 78 -
Results show that the location of the overlap in the preform (interval versus no
interval) affects infusion time. The increase in infusion time is greater when more
than one overlap occurs at the same position in the thickness of the preform. As
expected, preforms including an overlap present longer infusion times on their lower
face since the overlap results in an increase of the thickness of the preform.
When compared to an 8-layer preform with no overlap, lead-lag is 2.3 times longer
for preforms presenting an overlap without interval and 1.6 times longer for
preforms presenting an overlap with interval (59 mm after 2000 sec, upper flow
front 600 mm away from the inlet).
Overlapping fabric layers in a preform completely modifies the impregnation
kinetics. Trans-plane flow still leads the impregnation but since thickness changes
occur and the upper flow front is ahead of the lower flow front, the area on each
side of the overlap is filled prior to the area of increased thickness as shown in
Figure 3.33. Consequently a dry patch occurs in the preform, impregnated by in-
plane flow occurring straight behind the flow front.
Fig. 3.33. Impregnation of the preform when an overlap occurs. Lower side of the preform.
3.6. Conclusions
This experimental investigation of lead-lag reveals that three major phases occur
during preform impregnation. The first phase occurs at the onset of the infusion
where in-plane flow in both the distribution medium and the preform leads the
impregnation. Lead-lag increases due to the high difference in in-plane permeability
between the distribution medium and the reinforcement. Trans-plane flow has
already occurred. The length of the lead-lag developed during this phase determines
Flow direction
Overlap area
Impregnated fabric
Dry fabric
- 79 -
the lead-lag distance that characterises the rest of the infusion. The second phase
occurs when trans-plane flow reaches the lower side of the preform for the first time
and takes the lead on in-plane flow. Lead-lag reaches its maximum length. In the
meantime flow in the distribution medium remains in-plane, as does flow in the
preform behind the front. Lead-lag decreases progressively due to a shorter
distance to cover through the thickness of the preform compared to the in-plane
dimensions. The third phase occurs when lead-lag reaches a steady state. The latter
does not depend on preform thickness but does depend on the permeability
difference between the distribution medium and the reinforcement as it is directly
linked to the velocity of the flow in the distribution medium and the preform. The
exact orientation of the trans-plane flow vector cannot be determined. Yet it is
possible to determine it analytically, based on knowledge of permeability and
compaction of the stack.
The difference in in-plane permeability between the distribution medium and the
preform as well as the thickness and trans-plane permeability of the preform control
the occurence and length of the lead-lag. In addition the in-plane permeability of
both the distribution medium and the reinforcement is more affected by the
pressure gradient than the trans-plane permeability of the preform. However it is
not clear whether the distribution medium behaves as an open channel or a porous
media. Further investigation is required, which is the focus of the next chapter.
Lead-lag can be eliminated by cutting the distribution medium shorter than the
preform. In this case, in-plane flow takes the lead on the infusion. The distance
necessary to eliminate lead-lag does not depend on preform thickness, unlike time,
to recover in-plane flow. It is assumed that this distance is a function of preform
permeability.
Finally when thickness variations occur in a preform, it is recommended to infuse
the latter in the direction of decreasing thickness in order to reduce both lead-lag
and infusion time. In addition overlaps in preforms increase infusion time and result
in dry patches which are slowly impregnated by inplane flow coming behind the
main flow front.
- 80 -
References
1. W. H Seemann. Plastic transfer moulding techniques for the production of
fibre reinforced plastic structures. US patent No 4902215, filed March 1989.
2. H. M. Andersson, T.S. Lundstrom, B.R. Gebart, R. Langstrom. Flow
enhancing layers in the vacuum infusion process. Polymer composites, Oct
2002, 23, N
o
5, p. 895-901.
3. W.S Leenders, P. Marissen, L. Drift, B. Boon. Special problems due to
vacuum injection of large ship structures. Flow Processes in Composite
Materials conference, 1999, Plymouth, England, p. 329-336.
4. M. Thibaudeau, R.A. Shenoi. Understanding flow behaviour in the resin
infusion process. Tenth European Conference on Composite Materials (ECCM-
10), June 3-7, 2002, Brugge, Belgium.
5. T. Searle, J. Spooner, S. Grove, R. Cullen, S. Davy. Manufacture of large
composite structures by vacuum resin infusion. 24
th
International SAMPE
Europe conference, April 2003, Paris, France, p. 113-120.
6. M. Feiler, L. Ischtschuk. Vacuum Assisted Resin Infusion (VARI): on the way
to serial production. SAMPE conference, 2003, Paris, France, p. 683-691.
7. M. Andersson, S. Lundstrom, B.R. Gebart, R. Langstrom. Development of
guidelines for the vacuum infusion process. Fibre Reinforced Composite
Conference (FRC), 2000, p. 113-120.
8. X. Song, A.C. Loos, B. Grimsley, R. Cano, P. Hubert. Modelling manufacture
of textile composites by VARTM. 6
th
International Conference on Textile
Composites (TexComp 6), Sept 2002, Philadelphia, USA.
9. X. Sun, S. Li, L.J. Lee. Mold filling analysis in vacuum assisted resin transfer
molding. Part I: SCRIMP based on a high permeable medium. SPE Technical
journals, Polymer Composites, Dec. 1998, 19, N6, p. 807-817.
- 81 -
10. A. Hammami, R. Gauvin, F. Trochu. Modelling the edge effect in liquid
composites moulding. Composites Part A, 29, 1998, p. 603-609.
11. K. Han, S. Jians, C. Zhang, B. Wang. Flow modelling and simulation of
SCRIMP for composites manufacturing. Composites Part A, 31, 2000, p. 79-
86.
12. S.G. Advani et all. Flow and rheology in polymer composites manufacturing.
Newark, Elsevier, ISBN 0-444-89347-4 (Vol. 10), 1994. Chapter 12, p. 465-
515.
13. A. Ragondet, N. C. Correia, F. Robitaille, A. C. Long, C. D. Rudd.
Experimental investigation and modelling of the vacuum infusion process.
Tenth European Conference on Composite Materials (ECCM-10), June 3-7,
2002, Brugge, Belgium.
14. K-T. Hsiao, R. Mathur, S. G. Advani, J. W. Gillespie Jr, B. K. Fink. A closed
form solution for flow during the vacuum assisted resin transfer molding
process. Journal of Manufacturing Science and Engineering, Aug 2000, 122,
p. 463-475.
15. M. J. Tari, J-P. Imbert, M. Y. Lin, A. S. Lavine, H. T. Hahn. Analysis of resin
transfer moulding with high permeability layers. Journal of Manufacturing
Science and Engineering, Aug 1998, 120, p. 609-616.
- 82 -
Chapter 4
Experimental analysis and modelling of flow in the distribution
medium
4.1. Introduction
The use of a distribution medium in VI results in the occurrence of a lead-lag as
shown in Figure 4.1. Study of lead-lag in chapter 3 did not show what governs flow
in the distribution medium. This chapter aims to characterise flow behaviour
through this material. Both experimental and analytical studies were conducted. The
permeability of the distribution medium was measured and compared to CFD
(Computational Fluid Dynamics) models where a simplified geometry was used to
reproduce the architecture of the material. A series of VI experiments were
conducted to study the infusion of preforms covered by varying number of
distribution medium layers. The effect of number of distribution medium layers on
infusion time was analysed and conclusions about flow behaviour in the distribution
medium were made.
In addition the nesting of the material was modelled in order to determine its effect
on permeability. Nesting is characterised by a mis-alignement of the distribution
medium layers when using more than one layer of material. It is suspected to affect
the flow enhancement properties of the distribution medium.
4.2. Background
Flow phenomena in VI when using a distribution medium have already been studied
to analyse lead-lag during infusion and its effect on part quality (see chapter 2 and
3). Tari [1], studied the effect of flow layers on flow front. The author describes the
creation of a transverse flow due to the presence of the distribution medium and
mentions that the use of such a fabric reduces infusion time by 87%. Though
neither the type and permeability of both the preform and the distribution medium
are mentioned. The analysis reveals that the fibre saturation from in-plane flow
becomes greater as the distribution medium becomes thinner.
- 83 -
a). No distribution medium used: no lead-lag
b). Use of a distribution medium on the upper side
of the preform: lead-lag occurs
Fig. 4.1. Effect of distribution medium on flow progression in VI.
Infusion of an E-glass chopped strand mat M5 (Vetrotex, 450 g.m
-2
, cured resin).
The reduction in thickness involves a reduced flow rate and it takes more time for
the resin to travel from the distribution medium to the preform, thus reducing lead-
lag. The author concludes that the apparent permeability of a lay-up which involves
a distribution medium is close to that of the distribution medium itself.
Andersson [2] studied experimentally the effect of the number of distribution
medium layers on infusion time. The reinforcement infused was a non-crimp
stitched bonded material (800 g.m
-2
) while the distribution medium layer consisted
of a chopped strand mat (300 g/m
-2
). The number of chopped strand mat layers on
top of a 8-layer preform of non-crimp stitched bonded reinforcement varied from 0
to 3 and its effect on infusion time was measured. Results showed that adding 1 to
2 layers of flow enhancing fabric decreased infusion time significantly. However
Flow front position on both upper
and lower side of the preform
Dry fabric Impregnated fabric
Flow front position on the lower
side of the preform
Flow front position on the upper
side of the preform
Impregnated fabric Dry fabric
- 84 -
additionnal time saved from the third layer was small. It was assumed that the
flexibility of the bag influenced the results.
Sun [3] measured the effect of the number of distribution medium layers on
distribution medium permeability. A knitted fabric with high porosity was used for
the experiments but no information was given on the nature of the material. The
number of layers was varied from 1 to 7. The authors assumed that the
permeability in 1 layer of material should be smaller due to the frictional force
caused by the wall of the mould. However results showed the opposite and the
permeability was found to be higher in the 1 layer case. This was attributed to the
nesting effect occuring when using several layers of material.
The effect of fibre nesting on permeability was studied by Papathanasiou [4] who
determined analytically the effect of random perturbations in unidirectionnal arrays
of cylindrical fibres on transversal permeability. The authors studied the variations
of permeability between a perfect square packing and random perturbations in the
location and size of the fibres. It was found that, for high porosities greater than
0.80, random perturbations in fibre size and - or location around averages
corresponding to a perfect square array have no effect on permeability. It was
concluded that at high levels of porosity the random placement of fibres leaves
large channels between groups of fibres, which results in an average lower flow
resistance. In addition it was found that a fully random structure showed a
permeability slightly higher than a perfect square array for porosities greater than
0.80.
Simacek [5] modelled the presence of a distribution medium in LIMS (Liquid
Injection Moulding Simulation), by adding two rows of mesh on top of the mesh
representing the reinforcement (LIMS is a flow modelling software developed by
the University of Delaware, which allows predicting flow through porous media
during RTM process). Results showed that modelling flow through the distribution
medium was not realistic when setting isotropic permeability. But setting a trans-
plane permeability two orders of magnitude lower than in-plane permeability gave
better results.
- 85 -
4.3. Assumptions
The aim of this study is to determine whether the distribution medium behaves like
a porous media or an open channel. In a porous media like material, the
permeability remains constant whatever the thickness of the material. Contrary to
the porous media, the permeability of an open channel vary with the size of the
channel. Therefore a good mean of determining flow behaviour through the
distribution medium is to measure its permeability for different number of layers. If
the permeability is constant with increasing layers number then the distribution
medium behaves as a porous media. However the distribution medium is expected
to nest with increasing layers number and its permeability to increase with the
number of layers. Therefore the distribution medium is assumed to behave as an
open channel. In addition the effect of nesting and walls distance on distribution
medium permeability, when increasing the number of material layers has to be
investigated.
4.4. Permeability measurement of the distribution medium
4.4.1. Materials
Permeability measurements were conducted on the HDPE chain stitched distribution
medium H8036 (see Appendix 2). Infusions were conducted between two non-
deformable flat plates (a steel base and a glass cover). Mineral oil HDX30 from
Trent Oils was used for the experiments. The viscosity of the oil was close to that of
the polyester resin 701 PAX: respectively 230 mPa.s and 177 mPa.s at 20C. It was
recorded prior to each experiment with a Brookfield LV-DII viscometer. A smooth
polyamide vacuum bag and sealant tape, identical to that used for VI experiments
(see Appendix 2), sealed the cavity around the plates. Aluminium shims located
between the two halves of the mould were used to adjust cavity thickness. They
were 5 mm wide and 700 mm long and were placed along the edges of the mould.
Infusions were conducted for 3 different cavity thicknesses: 0.998 mm, 1.960 mm
and 2.958 mm, which corresponded to the thickness of the shims.
4.4.2. Experimental set-up
The tooling consisted of a non-deformable steel plate and a 25 mm thick glass
cover. Distribution medium layers were laid on the steel base and covered with the
- 86 -
glass plate (see Figure 4.2). A polyamide bag sealed the cavity. An inlet connected
the mould cavity to the oil feeder and opposite to the inlet, an outlet connected to
an Edwards RV5 pump provided the air extraction. Both inlet and outlet used 6 mm
bore PVC pipe connections with 15 mm diameter steel springs to form full width
inlet and outlet galleries. The objective was to promote rectilinear flow throughout
the cavity.
Prior to permeability measurements, the thickness of the mould cavity was
measured under vacuum to ensure the non-bending of the steel base or the glass
cover. Measurements were conducted in moulding conditions under 101.3 mbar of
absolute pressure and for 0.958 mm, 1.960 mm and 2.958 mm cavity thickness.
The cavity thickness was measured around the periphery of the mould, on centre,
right and left side using a Vernier Caliper mounted on a steel support. Results
displayed in Table 4.1 show that thickness variation was very low.
Fig 4.2. Permeability measurements of the distribution medium.
Experimental set-up.
1, 2 and 3 layers of distribution medium were impregnated in respectively 0.998
mm, 1.960 mm and 2.958 mm thick cavity. Each experiment was repeated 3 times.
The tests were conducted for either weftwise or warpwise flow. The porosity of the
material was 0.89. It was determined using Equation 4.1.
Pump
Sealant tape
Vacuum bag
Glass plate
Steel base plate
Spring (inlet)
Spring (outlet)
- 87 -
f
de
h
S n
p
c

= 1 4.1
Where c is the porosity, n the number of material layers, S
de
the surface density of
the material, h the thickness of the lay-up and p
f
the density of the material.
Figure 4.3 shows the architecture of the distribution medium where the stitch and
inserted filaments appear in black and green respectively.
Fig. 4.3. Architecture of the distribution medium.
Table 4.1. Measurement of cavity thickness prior to experiments.
Thickness
theory
(mm)
Average
measured
thickness (mm)
Maximum
measured
thickness
(mm)
Minimum
measured
thickness (mm)
Deviation on
maximum
thickness (%)
Deviation on
minimum
thickness (%)
0.998 1.021 1.079 0.964 5.3 5.9
1.960 1.953 2.019 1.902 3.3 2.6
2.958 2.974 3.069 2.884 3.1 3.1
Fill times were recorded for each experiment and the permeability was calculated
using an integration of Darcys law (see Equation 4.2).
P
x
K
t
A

=
2
2
c u
4.2
Pressure losses in the inlet pipe were deducted from the initial vacuum level set.
They were calculated as follow:
Pressure loss due to gravity
Warp
Weft
Stitch
Filament
- 88 -
z g P
g
= A p 4.3
With z being the height separating one end of the pipe, in the resin bucket, from the
other end of the pipe, at the inlet of the mould. p is the density of the fluid and g
the acceleration.
Pressure loss due to pipe wall
2
8
p
w
r
x v
P
u A
= A 4.4
v is the velocity of the fluid in the pipe, r
p
the radius of the pipe, u the viscosity of
the fluid.
The pressure value used for permeability calculation was then:
w g
P P P A A = A 4.5
Tables 4.2 and 4.3 shows the values of pressure losses in the pipe for all the tests
conducted.
Table 4.2. Pressure losses in pipe. Weftwise flow.
P
g
(Pa) P
w
(Pa) Cavity thickness
(mm)
Test 1 Test 2 Test 3 Test 1 Test 2 Test 3
0.998 9466 9466 9466 4968 4948 4919
1.960 9466 9466 9466 4790 4563 4415
2.958 9466 9466 9466 4356 4375 4405
Table 4.3. Pressure losses in pipe. Warpwise flow.
P
g
(Pa) P
w
(Pa) Cavity thickness
(mm) Test 1 Test 2 Test 3 Test 1 Test 2 Test 3
0.998 9466 9466 9466 4819 4780 4691
1.960 9466 9466 9466 4701 4632 4484
2.958 9466 9466 9466 4602 4662 4662
4.4.3. Results
Table 4.4 shows both weftwise and warpwise permeability values of the distribution
medium. Firstly the orientation of the distribution medium to the direction of the
flow does not affect considerably the permeability of the material. Secondly
- 89 -
permeability increases with cavity thickness. Still, the porosity is supposed to be
constant as the ratio of cavity thickness to the number of distribution medium layers
is constant. This is due to the nesting of the fabric when using more than one layer
of distribution medium. Consequently the porosity of the distribution medium varies
with the number of layers and so does the permeability. Finally results are very
repeatable. This shows that, although nesting occurs when using more than one
layer of distribution medium, the size of the flow paths created in and between
fabric layers does not vary significantly from one test to another. Consequently flow
through the distribution medium can be compared to flow of a fluid through an open
channel.
Table 4.4. Experimental permeability of the distribution medium
Average K (10
-8
m
2
) Average K (10
-8
m
2
)
Cavity thickness (10
-3
m)
Weftwise Warpwise
0.998 1.26 0.02 1.06 0.01
1.960 2.27 0.08 2.36 0.06
2.958 2.69 0.07 2.76 0.02
4.5. 2D CFD analysis
The next step of the study focussed on the modelling of flow through the
distribution medium using CFD analysis software Fluent
TM
. Cases similar to that used
in permeability measurements were modelled. Fluent
TM
allows the modelling of fluid
flow and heat transfer in complex geometries. For both two-dimensional and three-
dimensional study, geometry and mesh generation were carried out using Gambit
TM
,
Fluents
TM
geometry and mesh generation software. Simulations were run using a
steady-state flow condition, incompressible fluid, laminar flow and with a
convergence criterion of 1.10
-5
. The velocity distribution was obtained from Fluent
TM
simulation and the permeability was calculated using Darcys law:
P
x
v K
A
A
= u c 4.6
u was set to 0.23 Pa.s (viscosity of the oil HDX 30 at 20C).
4.5.1. Geometry
A simplistic approach consisting of arrays of parallel cylinders was considered. The
architecture of the distribution medium was approximated using Gambit
TM
according
to the orientation of the mesh pattern, the number of layers and the thickness of
- 90 -
the cavity. Figure 4.4 shows the geometry modelled. The properties of the fluid
were that of the oil used for permeability measurements (density, viscosity).
Boundary conditions were as follows: imposed pressure at the inlet and vent at the
outlet so that the pressure difference was identical to that used in experiments. The
geometry was designed between two parallel flat plates set as solid walls. The
average flow velocity value obtained from the flow simulation was used to
determine the permeability of the model. The analysis was conducted for all cases
studied experimentally: 1 layer of distribution medium in a 0.998 mm thick cavity, 2
layers in a 1.960 mm thick cavity and 3 layers in a 2.958 mm thick cavity. Flow
simulations were run for both weftwise and warpwise flow. Models were 10.00 mm
long in the weftwise geometry and 11.40 mm long in the warpwise geometry. The
mesh was made of 1000 triangular elements and the nodes were separated by 0.1
mm from each other. The effect of the accuracy of the mesh on the results was not
deeply analysed as the number of elements per mesh was limited to 1000 - 3000
due to the limited running capacity of the computer. But, one of Gambit functions,
which allows obtaining an optimal mesh distribution, was used for all the meshes
generated for this CFD study. The porosity was constant and equal to 0.89 for all
geometries. Table 4.5 shows the cases modelled.
Fig 4.4. Geometry used in the 2D CFD analysis.

]
] ]
]
Weftwise flow
Weftwise flow
Warpwise flow
Warpwise
flow
- 91 -
Table 4.5. 2D CFD models characteristics
Case Flow direction Cavity thickness Number of layers
1 0.998 1
2 1.960 2
3
Weftwise
2.958 3
4 0.998 1
5 1.960 2
6
Warpwise
2.958 3
4.5.2. Results
Figures 4.5 and 4.6 show the pressure and velocity distribution for either weftwise
or warpwise flow. High pressure velocity areas are represented in red, low pressure
and velocities are reprented in blue.
Weftwise flow
Pressure distribution (Fluent
TM
) Velocity profile (Fluent
TM
)
Average
velocity (m/s)
0.011
0.027
0.038
Fig. 4.5. Pressure and velocity distribution. 2D CFD model. Weftwise flow. Cases 1 to 3.
Warpwise flow
Pressure distribution (Fluent
TM
) Velocity profile (Fluent
TM
)
Average
velocity (m/s)
0.017
0.029
0.034
Fig. 4.6. Pressure and velocity distribution. 2D CFD model.
Warpwise flow. Cases 4 to 6.
Pressure losses occur in the vicinity of the filaments. The permeability of the models
follows the same tendency than that obtained experimentally, i.e the permeability
increases with cavity thickness and for constant porosity (see Figure 4.7). However,
analytical values are higher than experimental ones due to an oversimplification of
- 92 -
the filaments geometry. The permeability with warpwise flow shows closer
agreement with experiments.
Fig. 4.7. Permeability of the distribution medium.
Comparison between experiments and 2D CFD models.
These models represent an array of parallel filaments whereas in reality, the chain
stitch pattern results in both weftwise and warpwise oriented filaments in the
distribution medium (see Figure 4.4). The next step of the study consisted of
modelling the distribution medium architecture in three dimensions so both weft and
warp directions were modelled in a single geometry.
4.6. 3D CFD analysis
4.6.1. Geometry
The three-dimensional CFD analysis was modelled as an overlaying of the two 2D
geometries perpendicular to each other so no interlacing occurred (different from
the real case) as shown in Figure 4.8. Boundary conditions and material properties
were identical to the 2D analysis. Again both weftwise and warpwise flows were
modelled. According to the direction of flow the pressure was imposed at one edge
of the model and the opposite edge was set as a vent (see Figures 4.9 and 4.10).
The other edges of the model were set as walls to replicate the mould and glass
cover surfaces used in the experiments. The meshes were made of triangular
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
8.0
0.0 1.0 2.0 3.0
Cavity thickness (mm)
P
e
r
m
e
a
b
i
l
i
t
y
(
1
0
-
8
m
2
)
2D Weftwise
Exp Weftwise
2D Warpwise
Exp Warpwise
- 93 -
elements (1000 to 3000 elements depending on the size of the model) and the
nodes were separated by 0.1 mm from each other.
Fig. 4.8. 3D CFD model. Architecture of the distribution medium, 10 mm by 11.4 mm. 1
(0.998 mm thick), 2 (1.960 mm) and 3 layers (2.958 mm).
4.6.2. Results
Figures 4.9 and 4.10 show pressure and velocity distributions for flow through 1
layer of distribution medium. The distribution medium is oriented respectively
weftwise and warpwise to the flow direction. Results for flow through 2 and 3 layers
of distribution medium are shown in Figure 4.11 and 4.12.
Pressure distribution (Fluent
TM
) Velocity profile (Fluent
TM
)
Fig. 4.9. Pressure and velocity distribution. 1 to 3 layers. Weftwise flow.
Table 4.6 shows the corresponding velocity values. Figure 4.13 shows the
corresponding permeability values and compares them to experimental results. 3D
models show different results according to the orientation of the flow compared to
1 Layer 3 Layers
2 Layers
Inlet
Outlet
Inlet
Outlet
- 94 -
that of the stitch. Warpwise permeabilities are very close to experimental ones
whereas weftwise ones remain higher. This is similar to what was obtained in 2D.
Pressure distribution (Fluent
TM
) Velocity profile (Fluent
TM
)
Fig. 4.10. Pressure and velocity distribution. 1 to 3 layers. Warpwise flow.
Table 4.6. 3D CFD velocity values in 1, 2 and 3 layers
of fabric and for weftwise and warpwise flow
Weftwise flow
Number of layers 1 2 3
Velocity (m/s) 0.0038 0.0179 0.0202
Warpwise flow
Number of layers 1 2 3
Velocity (m/s) 0.0098 0.139 0.0163
Pressure distribution (Fluent
TM
) Velocity profile (Fluent
TM
)
Fig. 4.11. Pressure and velocity distribution. 1 to 3 layers. Warpwise flow.
Inlet
Outlet
Inlet
Outlet
Inlet Inlet
Outlet Outlet
- 95 -
Pressure distribution (Fluent
TM
) Velocity profile (Fluent
TM
)
Fig. 4.12. Pressure and velocity distribution. 1 to 3 layers. Warpwise flow.
Fig. 4.13. Permeability of the distribution medium.
Comparison between experiments, 2D and 3D CFD analysis.
4.6.3. Modelling the nesting effect
The effect of nesting on distribution medium permeability was modelled. Nesting
was represented by moving the position of the filaments and layers in both x and y
directions. x represented the distance between inlet and outlet (flow direction) and
y the thickness of the cavity. 2 to 3 layers of distribution medium were modelled in
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
8.0
0.0 1.0 2.0 3.0
Cavity thickness (mm)
P
e
r
m
e
a
b
i
l
i
t
y
(
1
0
-
8
m
2
)
2D Weftwise
3D Weftwise
Exp Weftwise
2D Warpwise
3D Warpwise
Exp Warpwise
Inlet Inlet
Outlet Outlet
Inlet
Outlet Outlet
Inlet
- 96 -
2D. Figure 4.14 shows examples of geometries for the 2 layers models. Flow was
oriented warpwise. Results from the 3 layer models are displayed in Figures 4.15 to
4.17.
Case 1
Case 2
Fig. 4.14. Modelling the nesting of the distribution medium on fluid velocity (Fluent
TM
).
Cases presented here: two layers of distribution medium.
Table 4.7. Velocity through 2D CFD model of distribution medium.
Average velocity (m/s)
Case 1 Case 2
0.0176 0.0238
0.0180 0.0181
0.0091 0.0089
Case 3 Average velocity (m/s)
0.0211
0.0199
0.0150
Fig. 4.15. Modelling the nesting of the distribution medium on fluid velocity (Fluent
TM
).
Cases presented here: three layers of distribution medium.
x
y
x
y
Outlet
Outlet
Inlet
Inlet
x
y
- 97 -
Case 4 Average velocity (m/s)
0.0221
0.0211
0.0177
Fig. 4.16. Modelling the nesting of the distribution medium on fluid velocity (Fluent
TM
).
Cases presented here: three layers of distribution medium.
Case 5 Average velocity (m/s)
0.0300
0.0227
0.0230
Fig. 4.17. Modelling the nesting of the distribution medium on fluid velocity (Fluent
TM
).
Cases presented here: three layers of distribution medium.
Figure 4.18 shows the corresponding permeability values obtained and compares
them to experimental results and 2D simulations where no nesting occured. Results
show that the position of the filaments affects significantly the permeability of the
model. However, experimental measurements reveal good repeatability and no very
significant permeability changes. In reality, the distribution medium layers always
nest which explains why permeability measurements were repeatable and why
experimental permeability values are smaller than analytical ones. But, nesting does
not occur in a single layer of distribution medium since the filaments cannot move
easily because of the knitted pattern. A more realistic explanation is that since
experimental permeability measurements were very repeatable, another factor,
different from nesting, affects the permeability of the distribution medium more
x
y
x
y
- 98 -
significantly and tends to mask the effect of nesting. This parameter is suspected to
be the walls of the mould cavity, or in VI the vacuum bag. Still, this assumption has
to be investigated and validated.
Fig. 4.18. Permeability of the distribution medium as a function of cavity thickness.
Comparison between experiments, 2D CFD analysis with no nesting and with nesting.
Warpwise flow only.
4.7. Experimental analysis of the distribution medium
Flow through the distribution medium was analysed experimentally during VI. The
study consisted of infusing preforms of constant thickness and varying the number
of distribution medium layers on top. Then infusion times could be compared and
sections of the distribution medium could be analysed microscopically to quantify
nesting and vacuum bag conformation and measure their effect on flow properties.
4.7.1. Materials
A triaxial E-glass stitched fabric E-TLX 1169 (+ 45 / 0 / - 45, 1169 g/m
2
) from
Vetrotex was used for the experiments (see Appendix 1 and 2). The distribution
medium consisted of a Palmhive Textiles HDPE warp knitted chain stitched pattern
fabric with inserted filaments. The materials also included a woven nylon bleed out
fabric used to release other bagging materials from the cured part and a smooth
nylon vacuum bag. The resin system was a pre-accelerated isophthalic polyester
resin 701 PAX from Scott Bader initiated with 1% by mass of Butanox M50 from
Akzo Chemie. Technical details of the materials are available in Appendix 2.
Warpwise - no nesting (CFD modelling)
Warpwise nesting (CFD modelling)
Warpwise experimental
0.00E+00
1.00E+00
2.00E+00
3.00E+00
4.00E+00
5.00E+00
6.00E+00
7.00E+00
0.5 1.5 2.5 3.5
Cavity thickness (mm)
P
e
r
m
e
a
b
i
l
i
t
y
(
1
.
1
0
-
8
m
2
)
Perpendicular
Experimental perpendicular
(
1
0
-
8
m
2
)
7.00
6.00
5.00
4.00
3.00
2.00
1.00
0.00
- 99 -
4.7.2. Experimental set-up
9-layer preforms (0.2 m by 0.7 m) were placed on a 1 m by 1 m glass plate and
covered with a bleed out fabric. Layers of distribution medium, oriented weftwise to
the direction of flow, were added on top of the stack. Their number varied from 1 to
5. The inlet was connected to a 2 litre plastic bucket containing accelerated resin.
The catalyst was added immediately before infusion. The outlet was connected to an
Edwards RV5 pump. A resin trap was placed between the outlet and the pump in
order to store spare resin flowing out of the mould. Both inlet and outlet used PVC
pipe connections (12 mm external diameter, 3 mm wall thickness) and acetal resin
elbow connectors (12 mm diameter). 15 mm diameter steel springs were used on
both sides to promote rectilinear flow throughout the cavity and to keep uniform
vacuum distribution over the outlet side. The vacuum level was 101.3 mbar of
absolute pressure, i.e the pressure difference driving the flow was of 900 mbar.
Infusion times were recorded for constant flow front position intervals, on both
upper and lower faces of the preform. Recordings were made on the centre line,
every 50 mm, with 0 mm corresponding to the inlet and 700 mm to the outlet. The
experimental set-up is shown in Figure 4.19 and is similar to the set-up used for
experiments described in Chapter 3.
Fig. 4.19. Experimental set-up
Resin trap
Vacuum pump
Spring (outlet)
Glass plate
Fabric and bagging materials
Spring (inlet)
- 100 -
4.7.3. Results
Figure 4.20 shows infusion time on the upper side of the preform as a function of
the number of distribution medium layers. It expresses the time needed for flow to
progress from 0 mm to 700 mm between the inlet and outlet of the mould cavity.
As expected each additional layer of distribution medium decreases infusion time
further; but infusion time is not exactly inversely proportional to the number of
distribution medium layers, i.e the percentage of time saved with each additional
layer decreases with the number of layers.
This phenomenon is directly due to the nesting of the distribution medium layers.
Figure 4.21 shows that the stitched chains of the different layers do not align on
each other resulting in an increase of the thickness, which is not proportional to the
number of distribution medium layers (see Figure 4.22). A direct consequence is a
decrease of porosity in the distribution medium with the increasing number of
distribution medium layers (see Figure 4.23).
Fig 4.20. Effect of number of distribution mediumlayers on infusion time.
The trendline is a polynomial curve fit.
The percentage of time saved with each additionnal layer decreases with the number of layers.
Variations in porosity, determined from minimum and maximum thickness of the
flow layers (see Figure 4.23), are non negligible and higher for lower numbers of
distribution medium layers. This tendency is due to the bending of the polyamide
bag under vacuum, between the inserted filaments of the distribution medium (see
0
500
1000
1500
2000
2500
3000
0 1 2 3 4 5 6
Number of distribution medium layers
I
n
f
u
s
i
o
n
t
i
m
e
(
s
)
- 101 -
Fig 4.20. Effect of nesting and vacuum bag compliance
on the geometry of the distribution medium.
1 layer
2 layers
3 layers
4 layers
Bleed out fabric
Reinforcement Filament
Stitch
Nesting Vac. Bag compliance
5 layers
Vacuum bag
0.5 mm
0.5 mm
0.5 mm
0.5 mm
0.5 mm
- 102 -
Figure 4.21). This phenomenon affects more low numbers of distribution medium
layers as the loss in thickness, due to bag deformation, is relatively more important,
compared to the total thickness of the distribution medium (see Figure 4.23).
The thickness of the distribution medium was measured under an optical
microscope. Sections of the different laminates were cut, cast into polyester resin
and polished (3 samples per laminate). 20 readings of the thickness of the
distribution medium were taken per sample. The minimum thickness of the
distribution medium corresponded to areas where the bag was bending (see Figure
4.21). The maximum thickness of the distribution medium corresponded to the
areas where the bag was in contact with the stitch of the distribution medium (see
Figure 4.21). The values displayed in Figure 4.22 show an average thickness,
minimum thickness and maximum thickness of the distribution medium as a
function of the number of distribution medium layers.
These values of minimum and maximum thickness were used to determine the
minimum, maximum and average porosity of the distribution medium. The porosity
Fig 4.21. Evolution of distribution medium thickness
with number of distribution medium layers. The trendline is a linear curve fit.
The total thickness of the distribution medium does not increase proportionnally with the number
of distribution mediumlayers.
0.0
0.5
1.0
1.5
2.0
2.5
0 1 2 3 4 5 6
Number of distribution medium layers
D
i
s
t
r
i
b
u
t
i
o
n
m
e
d
i
u
m
t
h
i
c
k
n
e
s
s
(
m
m
)
- 103 -
Fig 4.22. Evolution of distribution mediun porosity with number of distribution
medium layers. The trendline is a polynomial curve fit.
A consequence of the non-propotional increase of the distribution medium thickness with the
number of layers is a decrease of the porosity with increasing number of distribution medium
layers. In addition the effect of vacuum bag bending is seen through the difference between
maximum and minimum porosity in a given distribution medium lay-up. The effect of bag
bending on lay-up porosity is more significant for low number of distribution medium layers.
was determined from Equation 4.1 which describes the porosity as a function of the
surface density of the material, the number of material layers, the thickness of the
laminate and the density of the material.
4.7.4. Comments
Increasing the number of distribution medium layers used in VI results in an
increase of the distribution medium stack in-plane permeability. However, the
porosity of the stack decreases for an increasing number of distribution medium
layers. This is uncharacteristic of a porous media where the permeability is not a
function of the number of fabric layers but increases with porosity. These results
show that the porosity of the distribution medium and the number of distribution
medium layers are linked. Due to the nesting of the distribution medium layers, the
size and number of flow paths created in the distribution medium varies with the
number of layers. This implies substantial changes in distribution medium
permeability. In addition bag bending affects more significantly the permeability of
the distribution medium when low number of material layers are used.
60
64
68
72
76
80
84
88
92
0 1 2 3 4 5 6
Number of distribution medium layers
P
o
r
o
s
i
t
y
(
%
)
Nesting effect
Bag bending effect
- 104 -
4.8. Conclusions
This experimental and analytical study of flow through the distribution medium
shows that considering this material as an open channel is realistic. CFD modelling
analysis shows that it is possible to replicate the flow behaviour through this
material using simple geometries. 3D models show more accurate results than 2D
one as they represent more realistically the structure of the material. The 3D
warpwise flow model gives encouraging results and shows permeability values close
to experimental ones. In addition the study shows that, even if the distribution
medium layers nest when using more than one layer, flow behaviour is influenced
more by the conformation of the bag under vacuum than by the nesting itself. But
experimental analysis shows that the effect of the vacuum bag conformation affects
more a case with a low number of distribution medium layers.
Finally this study opens possibilities for the development of new flow enhancement
media where the actual architecture of the distribution medium can be replaced by a
network of open channels. Following this perspective the next Chapter proposes a
design methodology for the development of a distribution medium where the flow
enhancement is created by an array of channels separated by rhomboid cells.
- 105 -
References
1. M. J Tari., J-P Imbert, M. Y Lin, A. S Lavine, H. T Hahn. Analysis of resin
transfer moulding with high permeability layers. Journal of Manufacturing
Science and Engineering. Aug 1998, 120, p. 609-616.
2. P. Simacek, J. Lawrence, S. Advani. Numerical mold filling simulations of
liquid composite molding processes. Applications and current issues. 23
th
International SAMPE Europe Conference, April 2002, Paris, France, p. 137
148.
3. M. Andersson, S. Lundstrom, B.R. Gebart, R. Langstrom. Development of
guidelines for the vacuum infusion process. Fibre Reinforced Composite
Conference (FRC), 2000, p. 113-120.
4. X. Sun, S. Li, L.J. Lee. Mold filling analysis in vacuum assisted resin transfer
molding. Part I: SCRIMP based on a high permeable medium. SPE Technical
journals, Polymer Composites, Dec. 1998, 19, N6, p. 807-817.
5. T. D. Papathanasiou, P. D. Lee. Morphological effects on the transverse
permeability of arrays of aligned fibres. Polymer Composites, April 1997, 18,
No 2, p. 242-253.
- 106 -
Chapter 5
Novel vacuum membrane design methodology
5.1. Introduction
The numerous consumables used in vacuum infusion considerably increase the
purchase and labour costs. They all have a specific role in and, for example, are
used to improve part release (bleed out fabric), or reduce infusion time (distribution
medium). Many plastic pipes, valves and springs are also necessary to set-up the
infusion. In addition some consumables, such as the distribution medium, absorb
resin which further increases the production costs. A solution to decrease VI cost
consists of reducing the amount of consumables involved in the process. Previous
analysis, described in Chapter 2, shows that, for example, textured vacuum bags
have been developed to replace the use of a distribution medium. However,
experimental analysis demonstrates that this type of material does not successfully
replicate flow through the distribution medium.
5.2. Hypothesis
Previous analysis shows that the chain-stitched structure of the distribution medium
H8036 can be modelled as an open channel (see Chapter 4). The study of lead-lag
in VI, described in Chapter 3, shows that the latter is a direct function of preform
and distribution medium permeability. Yet, in VI the preform is subjected to a non-
homogeneous compaction due to the geometry of the part, which affects its
permeability and therefore the lead-lag. The structure of any distribution medium
currently used in VI is uniform on all the surface of the mould and does not allow
any adjustment of its permeability according to part geometry. A good means of
doing so, and thus increasing flow control in VI, involves the adaptation of the
permeability of the distribution medium to that of the preform so that the
permeability ratio between the two remains constant. This requires a new design of
the distribution medium where the structure of the material complements that of
the preform. One solution is to combine the flow enhancement properties of the
distribution medium with the flexibility and air-tightness of the vacuum bag in an
innovative vaccum membrane so that a separate distribution medium can be
eliminated. Therefore a new architecture can be envisaged, which favours low resin
- 107 -
consumption, permeability variability as well as better flow control, higher
repeatability of the infusion and increased part quality if combined with re-usable
materials. In addition, infusion strategies can be simplified.
This chapter proposes a methodology to design and develop such a vacuum
membrane. The latter consists of an array of channels separated by rhomboid cells.
The plan combines a computational and an experimental approach. The overall
objective is to develop a systematic analysis that defines the optimum geometry for
a membrane according to the flow properties required. Therefore the vacuum
membrane is peculiar to each component, facilitates flow control in VI and
contributes to improve efficiency and repeatability of the process. Results from
preliminary study are presented and a methodology for further development is
proposed.
5.3. Background
Innovative vacuum bags dedicated to the vacuum infusion process have already
been developed in the past, mainly motivated by a reduction of labour cost and
cycle time. The use of re-usable bags such as silicone membranes has been
reported since the late seventies [2] and are popular for their tear resistance, re-
usability and large percentage of elongation which helps the membrane to conform
to the preform geometry.
Williams [2] reports the development of a patented embosssed bag which eliminates
the need for a breather cloth. Such innovative vacuum bag has been developed by
Heider [3]. The author designed a double vacuum bag system called FASTRAC. The
aim was to reduce labour and waste by reducing the amount of consumables but
also to control and optimise resin flow during infusion with the objective of reducing
infusion time. The double vacuum bag, made of silicone, was made of a series of
parallel distribution channels in an outer bag while the inner bag was a flat
membrane. With this system no bleed out fabric and distribution medium were
required and the authors mention the significant reduction in cost and cycle time.
Still, although the vacuum bag was made of silicone and therefore re-usable, the
cost involved in the manufacture of the silicone bag was not mentioned, neither was
the resistance of the bag to styrene, assuming that polyester resin was used to
infuse the parts. The control of resin flow was made by adjusting the vacuum
- 108 -
between the inner and outer bag, thus adjusting the size of the flow channels.
Different FASTRAC configurations were analysed. First flow through a single channel
was tested then an array of parallel channels was used for experiments. Their effect
on resin flow was studied. Both channel width and depth were varied in order to
optimise resin flow. Results showed that flow progressed faster than in a
conventionnal vacuum infusion set-up. When an array of parallel channels was used
each channel developed its own flow and at some point, adjacent flow fronts
merged with the attendant risk of void formation. The configuration did not allow
permeability variation in the bag but provided a uniform flow path size over the
surface of the mould.
Although void formation in an array of channels, specifically dedicated to vacuum
membranes, has not been adressed widely, air entrapment due to the differences
between inter- and intra-tow permeability in preforms has been studied and can be
used as a reference for this study.
Parnas [3, 4] developed a model for flow of fluid during RTM process. The model
considers two competing flows, one macroscopic and one microscopic, which occurs
simultaneously, respectively in the preform and in individual fibre bundle that make
the preform, during the impregnation process. The one-dimensionnal model predicts
air entrapments in the fibre bundle and their evolution in size and location with flow
progression. The model assumes that the space within a fibre bundle is much less
than the space between the fibre bundles. Therefore permeability inside the bundle
is lower than outside the fibre bundle. It is assumed that, when the flow encounters
a fibre bundle, the flow splits into two parts and recoalesces on the downstream
side of the bundle before all the air has been expelled from the bundle by the
impregnation process.
Following Parnas model Sadiq [5] studied experimentally formation of voids inside
bundles. Results showed that lower volume fractions inside fibre bundles produced
smaller voids, as well as higher permeabilities. Similarly, Spaid [6] studied
experimentally transversal flow through square arrays of circular tows and found
that the size of the voids increased with the increasing porosity of the array. At high
porosities, the void encompassed nearly the entire tow.
- 109 -
Like Parnas, Pillai [7] modelled flow through dual scale porous media where the high
permeability inter-tow channels and the low permeability intra-tow regions are
represented by an array of parallel interspersed high and low porosity regions,
aligned with the flow direction.
Pearce [8] studied the effect of convergent flow front in RTM on air entrapment.
Contrary to Parnas [3] who neglected capillary forces, Pearce, like Breard [9],
assumed that capillary flow occured in the fibre bundle while channel flow lead
between the fibre bundles. Depending on the injection pressure, one lead the other.
At high pressure, channel flow lead on capillary flow and resulted in void formation
in the bundle due to a split and recoalescence of channel flow respectively at the
upstream and downstream side of the bundle, leaving an air pocket in the bundle
impregnated at lower speed. Experiments involving different point injection
strategies, two points mid-side ports, three points along one edge and one gallery
along one edge with mid-side port point, were conducted. Results showed that an
increased voidage occurred in the areas where flow from the different injection ports
met.
5.4. Preliminary study
5.4.1. Vacuum membrane geometry
CFD analysis was used to study an optimal vacuum membrane geometry that
combined the flexibility of the vacuum bag and the high permeability of the
distribution medium. The objective was to analyse flow and void formation. Analysis
of the distribution medium, described in Chapter 4, revealed that the latter can be
approximated by an open channel. Following this conclusion, a novel membrane
representing a network of open channels separated by rhomboids is proposed, as
shown in Figure 5.1.
The distribution medium and the classic soft vacuum bag were replaced by a single
bag, possibly made of re-usable material with cells embedded on one surface. The
rhomboid geometry was chosen for practical reason: a number of parameters, such
as cell size (side length), cell angle and channel size (height and width) could be
modified which eased the study of fluid flow and void formation.
- 110 -
a). Current set-up
b). Proposed set-up
Fig. 5.1. Proposed geometry to replace the present bagging arrangement.
5.4.2. Modelling
Analysis of the permeability through rhomboid cells
Rhomboid cells were modelled using Computationnal Fluid Dynamics (CFD) in order
to study the effect of both channel and cell size on membrane permeability. Figure
5.2 shows the geometry considered. The height of the channel was fixed at 1 mm to
minimise surface defects on the laminate. In addition the angle of the cell was set to
90.
Only the membrane geometry was modelled, the presence of a reinforcement
underneath the cells and channels was not considered. A ratio between channel
width and cell size, b/c (see Figure 5.2), was defined for each case studied. Three
different cases were modelled. Technical details are given in Table 5.1.
New geometry
Distribution medium
Vacuum bag
- 111 -
Amongst the geometries studied, one was defined such that the combined
dimensions of the diamond cells and open channels presented the same porosity as
in the current distribution medium so that both permeabilities could be compared.
The value of the ratio b/c was determined by calculating the ratio between the
surface occupied by the cell compared to the surface occupied by the channel. The
latter was assumed to be 0.89 to match the distribution medium porosity,
considering that the geometry was a square of 1 by 1 (dimensionless). The
corresponding ratio b/c was 2. The two other geometries modelled had a ratio b/c of
1 and 0.25.
Fig. 5.2. Unit cell of the reference geometry.
Geometry
Cell
Channel
Cell size
1/2 channel width in a unit cell
Unit cell
Smallest representation of
the unit cell
a
a/2
b
c
- 112 -
Table 5.1. Technical data, CFD analysis.
Parameter b/c = 2 b/c = 1 b/c = 0.25
Cell shape Rhomboid
Cell angle (
o
) 90
Cell size (mm) 2 2 4
Channel width (mm) 4 2 1
Channel height (mm) 1
Geometry dimensions (mm) 25.40 * 25.40 * 1 16.71 * 16.71 * 1 21.21 * 21.21 * 1
Porosity 0.89 0.89 0.36
Element shape Tetra
Element size in cell surface 0.2
Element size in volume 0.6
Pressure inlet (Pa) 2108.21 1858.74 2323.01
Fluid density (kg/m
3
) 1200.00
Fluid viscosity (Pa.s) 0.23
Flow velocity along Z axis (m/s) 0.0286 0.0219 0.0052
Permeability (10
-8
m
2
) 6.02 2.80 2.56
The geometries were modelled on Gambit and flow simulations were run on CFD
software Fluent 5 with the following boundary conditions: steady-state flow
condition, incompressible fluid, laminar flow and convergence criterion of 1.10
-5
.
Figure 5.3 shows one of the geometries modelled and the pressure gradient
obtained between the inlet and the outlet. Table 5.1 shows the boundary conditions
used for the flow simulation on Fluent 5, the size of the elements and the velocity
results. Figure 5.4 shows a typical representation of the velocity vectors during flow
simulation on Fluent (example with ratio b/c = 0.25). Velocity vectors are
represented by coloured arrows and the rhomboids by black cells.
Figure 5.3. Pressure gradient in a network of open channels and rhomboid cells.
Ratio b/c = 2 (Fluent 5).
- 113 -
Figure 5.5 shows the permeability obtained in the three cases modelled. It was
determined by implementing the velocity of flow in the channels, obtained from
Fluent 5, into Darcys law (see Equation 5.1). Results show that, as expected,
permeability increases with increasing channel size. The range of permeabilities
obtained is within the range of permeability of the distribution medium.
Fig. 5.4. Flow vectors in a network of open channels and rhomboids.
Ratio b/c = 0.25 (Fluent 5).
dx
dP K
v

=
u c
5.1
Fig. 5.5. Permeability as a function of the ratio b/c.
The trendline is a polynomial curve fit.
The geometry designed to match the permeability of the distribution medium shows
a higher permeability than expected: 6.02*10
-8
m
2
against 1.05*10
-8
m
2
and
X
Y
Z
Inlet
Cell
Channel
Outlet
Velocity
vectors
y = 0.3656x
2
+ 2.2937x
0.0
1.0
2.0
3.0
4.0
5.0
6.0
7.0
0 0.5 1 1.5 2 2.5
b/c
K
(
1
*
1
0
-
8
m
2
)
- 114 -
1.26*10
-8
m
2
for respectively warpwise and weftwise flow in the distribution
medium.
Incorporating the permeability values of the distribution medium in the polynomial
equation of Figure 5.5 shows that a ratio b/c of 0.48 to 0.55 should be used in the
cell geometry to match the permeability of the distribution medium. A ratio of 0.5
was then chosen and a new geometry was modelled. Table 5.2 shows the
dimensions and boundary conditions for the new geometry. The velocity obtained
from the flow simulation was 0.0161 m.sec
-1
which gave a permeability of 1.12*10
-8
m
2
for a porosity of 0.33. Figure 5.6 shows the velocity distribution in the new
geometry. The latter was taken as reference for the next step of the study which
was the analysis of void formation when impregnating a preform underneath a unit
rhomboid cell.
Table 5.2. Technical data, CFD analysis.
Parameter B/c = 0.5
Cell shape Diamond
Cell angle (
o
) 90
Cell size (mm) 2
Channel width (mm) 1
Channel height (mm) 1
Geometry dimensions (mm) 12.69 * 12.69 * 1
Porosity 0.33
Element shape Tetra
Element size in cell surface 0.2
Element size in volume 0.6
Pressure inlet (Pa) 1390.94
Fluid density (kg/m
3
) 1200.00
Fluid viscosity (Pa.s) 0.23
Flow velocity along Z axis (m/s) 0.0161
Permeability (10
-8
m
2
) 1.12
Analysis of flow in a unit cell
The reference geometry was taken as a model to vary dimensionnal parameters
(channel height and width, cell size, cell angle) and to look at their effect on channel
permeability and especially void creation. This part of the study was conducted on
LIMS (Liquid Injection Moulding Simulation) software and only unit cells were
modelled, as represented in Figures 5.7. The presence of the reinforcement
underneath the cell was also modelled and its permeability was varied to look at the
effect of permeability of the preform on void creation.
Indeed when the fluid flows through the channels, the reinforcement underneath the
cell is not immediately impregnated by the fluid due to permeability differences
- 115 -
between the channel and the preform. Therefore it is assumed that a void is created
underneath every cell and its size depends on the permeability of the channel and
the preform as well as on the surface occupied by the cell.
LIMS software allows void creation to be visualised during flow simulation.
However, when a void has been formed LIMS models it as a sink. Indeed the void
does not move with flow and the pressure inside the void is much lower than the
Fig. 5.6. Velocity distribution (m/s) in the reference geometry.
pressure outside the void so that its size decreases until it completely disappears.
Therefore it is not possible to predict its evolution, i.e size and location, with flow
front advancement.
The interest of such a study was to determine the limit for void formation and the
best geometry to be used to ensure no void occured according to the permeability
Centre = 0 mm
Wall = 0.5 mm
- 116 -
and the thickness of the preform. The models were designed using Hypermesh 4
and the flow simulations were run using LIMS.
Fig. 5.7. View of a unit cell as modelled in LIMS.
Results
Effect of the channel size on void formation
The analysis involved a 1 mm thick preform with in-plane permeability and trans-
plane permeability respectively of 1.0*10
-11
m
2
and 1.0*10
-12
m
2
(permeability of
the triaxial glass fabric EBX936, see Chapter 2). The analysis was conducted for 1
Unit cell
Rhomboid cell
Unit cell model
Channel
Rhomboid cell
Preform
- 117 -
mm, 2 mm and 4 mm wide channels with a permeability respectively of 4.3*10
-8
m
2
, 6.8*10
-8
m
2
and 9.1*10
-8
m
2
. The cell was 50 mm by 50 mm and the channel
was 1 mm high, as reported in Table 5.4. Figure 5.8 shows that a void always
occurs in the preform and its size increases with the size of the channel.
Table 5.4. Effect of channel size on void formation. Technical characteristics.
Case 1 Case 2 Case 3
Preform thickness (10
-3
m) 1
Preform permeability (m
2
)
K
x
= 1*10
-11
K
y
= 1*10
-11
K
z
= 1*10
-12
K
x
= 1*10
-11
K
y
= 1*10
-11
K
z
= 1*10
-12
K
x
= 1*10
-11
K
y
= 1*10
-11
K
z
= 1*10
-12
Preform V
f
0.5 0.5 0.5
Cell size (10
-3
m) 25 25 25
Cell angle () 90 90 90
Channel width (10
-3
m) 1 2 4
Channel permeability (10
-8
m
2
) 4.3 6.8 9.1
Channel height (10
-3
m) 1 1 1
Fluid viscosity (Pa.s) 0.23 0.23 0.23
Channel = 1 mm Channel = 2 mm Channel = 4 mm
Cell side
Preform side
Fig. 5.8. Flow modelling around unit cell using LIMS.
Effect of channel size on void formation.
This result is expected since the permeability of the channel increases with its size
whereas the permeability of the preform remains constant.
Inlet
Outlet
Wet
Dry
Void
Inlet
Outlet
- 118 -
Effect of the cell angle on void formation
This series of flow simulations involved variations of the cell angle. The channel
width was set to 1.0 mm, its thickness to 1.0 mm and the preform thickness was
1.0 mm. The in-plane permeability of the preform was 1.0*10
-11
m
2
and its trans-
plane permeability 1.0*10
-12
m
2
(see Table 5.5). The permeability of the channel
was 3.1*10
-8
m
2
. Results show that the size of the void increases when the angle of
the cell decreases in the flow direction (see Figure 5.9).
Table 5.5. Effect of cell angle on void formation. Technical characteristics.
Case 1 Case 2 Case 3
Preform thickness (10
-3
m) 1.0
Preform permeability (m
2
)
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
Preform V
f
0.5 0.5 0.5
Cell size (10
-3
m) 25 25 25
Cell angle () 90 67.5 45
Channel width (10
-3
m) 1 1 1
Channel permeability (10
-8
m
2
) 4.3 4.3 4.3
Channel height (10
-3
m) 1.0 1.0 1.0
Fluid viscosity (Pa.s) 0.23 0.23 0.23
Cell angle = 90 Cell angle = 67.5 Cell angle = 45
Cell side
Preform side
Fig. 5.9. Flow modelling around unit cell using LIMS.
Effect of cell angle on void formation.
This result is expected since a decrease in cell angle results in an increase of the
distance to cover underneath the cell before reaching the downstream side of the
cell where the two flows merge.
Cell angle
- 119 -
Effect of the cell size on void formation
For this series of simulations the channel width was set to 1.0 mm and its thickness
to 1.0 mm. The thickness of the preform was 1.0 mm, its in-plane permeability
1.0*10
-11
m
2
and its trans-plane permeability was 1.0*10
-12
m
2
(see Table 5.6). The
permeability of the channel was 3.1x10
-8
m
2
. The cell size varied from 25*10
-3
m to
30*10
-3
m and 40*10
-3
m. This size represents the length of one edge of the cell.
Results in Figure 5.10 show that a void occurs only underneath the 25 mm cell.
Table 5.6. Effect of cell size on void formation. Technical characteristics.
Case 1 Case 2 Case 3
Preform thickness (10
-3
m) 1.0
Preform permeability (m
2
)
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
Preform V
f
0.5 0.5 0.5
Cell size (10
-3
m) 25 30 40
Cell angle () 90 90 90
Channel width (10
-3
m) 1.0 1.0 1.0
Channel permeability (10
-8
m
2
) 4.3 4.3 4.3
Channel height (10
-3
m) 1.0 1.0 1.0
Fluid viscosity (Pa.s) 0.23 0.23 0.23
Cell size = 25 mm Cell size = 30 mm Cell size = 40 mm
Cell side
Preform side
Fig. 5.10. Flow modelling around unit cell using LIMS.
Effect of cell size on void formation.
- 120 -
This is explained by the fact that the permeability of the preform underneath the
cell is identical whatever the size of the cell. Because it takes a longer time for the
fluid to flow through longer channels in the larger cells, the preform has more time
to be impregnated before the two flows merge in the channels at the other end of
the cell.
Effect of preform permeability
Figure 5.11. shows the effect of preform permeability on void creation. The in-plane
permeability (K
x
, K
y
) and trans-plane permeability (K
z
) of the preform are given in
Table 5.7.
Table 5.7. Effect of preform permeability on void creation. Technical characteristics.
Case 1 Case 2 Case 3
Preform thickness (10
-3
m) 1.0
Preform permeability (m
2
)
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-10
K
y
= 1.0*10
-10
K
z
= 1.0*10
-11
K
x
= 1.0*10
-9
K
y
= 1.0*10
-9
K
z
= 1.0*10
-10
Preform V
f
0.5 0.5 0.5
Cell size (10
-3
m) 25 25 25
Cell angle () 90 90 90
Channel width (10
-3
m) 1.0 1.0 1.0
Channel permeability (10
-8
m
2
) 4.3 4.3 4.3
Channel height (10
-3
m) 1.0 1.0 1.0
Fluid viscosity (Pa.s) 0.23
K
x,y
= 1.10
-11
m
2
K
z
= 1.10
-12
m
2
K
x,y
= 1.10
-10
m
2
K
z
= 1.10
-11
m
2
K
x,y
= 1.10
-9
m
2
K
z
= 1.10
-10
m
2
Cell side
Preform side
Fig. 5.11. Flow modelling around unit cell using LIMS
TM
.
Effect of preform permeability on void formation.
- 121 -
The channel thickness was set to 1.0 mm and its width to 1.0 mm. The cell was 50
mm by 50 mm.
Results show that void size decreases when preform permeability increases and no
void occurs for a preform of in-plane permeability up to 10 times lower than the
permeability of the channel.
Effect of preform thickness
Simulations were performed for a 50 mm by 50 mm cell with 1.0 mm wide and thick
channel. The thickness of the preform was respectively 1.0 mm, 2.0 mm and 4.0
mm and its permeability 1.0*10
-11
m
2
in the in-plane directions and 1.0*10
-12
m
2
in
the trans-plane direction (see Table 5.8).
Table 5.8. Effect of preform thickness on void creation. Technical characteristics.
Case 1 Case 2 Case 3
Preform thickness (10
-3
m) 1.0 2.0 4.0
Preform permeability (m
2
)
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
K
x
= 1.0*10
-11
K
y
= 1.0*10
-11
K
z
= 1.0*10
-12
Preform V
f
0.5 0.5 0.5
Cell size (10
-3
m) 25 25 25
Cell angle () 90 90 90
Channel width (10
-3
m) 1 1 1
Channel permeability (10
-8
m
2
) 4.3 4.3 4.3
Channel height (10
-3
m) 1.0 1.0 1.0
Fluid viscosity (Pa.s) 0.23 0.23 0.23
Results reveal that, for a preform of permeability 1000 times lower than that of the
channels, voids occur when the preform presents the same thickness as the height
of the channel (see Figure 5.12). For thicker preforms no void appears on the lower
face of the preform. It could be expected that a void occured in every preform, not
necessarily visible when preform thickness increased. However a section of the unit
cell geometry in Figure 5.13 shows that no void occurs in the thicker preforms. This
can be explained by the presence of trans-plane flow in the preform which
decreases the velocity of the fluid in the bag channels and therefore prevents voids
from occuring before the two flows merge in the channels.
Summary
This preliminary modelling analysis shows that three main factors affect the
occurrence and size of voids underneath cells in the vacuum membrane. First the
- 122 -
difference in permeability between the channel and the preform, second the
difference in distance between the upstream and downstream side of the cell
Preform thickness = 1 mm Preform thickness = 2 mm Preform thickness = 4 mm
Cell side
Preform side
Fig. 5.12. Flow modelling around unit cell using LIMS.
Effect of preform thickness on void formation.
Fig. 5.13. Flow modelling around unit cell using LIMS.
Effect of preform thickness on void formation. View of a section through.
Void
Section
Direction of flow
No void
No void
- 123 -
whether considering the channel path or the straight path underneath the cell where
voids occur, third the thickness of the preform.
Figure 5.14 shows a summary of the results where the effect of the different
parameters analysed (channel size, cell size, cell angle, preform permeability) on
void size are expressed in term of the ratios of channel permeability over preform
permeability and the distance in a channel over the straight distance between
upstream and downstream side of a cell.
Fig. 5.14 Evolution of void size with channel and preform permeability as a function of the
distance to cover in the channel over underneath the cell between upstream and downstream
side of the cell.
Although the angle and size of the cells are significant factors that affect the size of
the voids, the permeability and thickness of the preform reported to the
permeability of the channels in the membrane are key factors that control the
occurrence of voids in the preform during infusion.
Practically, results show that a ratio of 10 between preform permeability and
channel permeability allows the obtaining of void free infusion considering that the
ratio between preform thickness and channel thickness is 1.
Additional flow simulations were conducted to analyse void creation in arrays of cells
and channels as shown in Figure 5.15 and 5.16. Table 5.9 shows the details of the
0
1000
2000
3000
4000
5000
6000
7000
8000
9000
10000
1.0 1.1 1.2 1.3 1.4 1.5
Distance in channel / Straight distance between merge points (m)
C
h
a
n
n
e
l
K
/
P
r
e
f
o
r
m
K
(
m
2
)
No void
Small void
Medium void
Large void
C
h
a
n
n
e
l
K
/
P
r
e
f
r
o
m
K
(
m
2
)
- 124 -
preform size and permeability as well as the dimensions of the cells and channels
and the boundary conditions used for the simulation.
The array was based on a unit cell previously modelled and where a void occurred
during flow simulation. Figures 5.15 and 5.16 show flow at different times during
the injection, respectively on upper side (cells side) and lower side (preform side) of
the stack.
Start of impregnation
End of impregnation
Fig. 5.15. Flow through an array of channels. Reference geometry. Upper side (cells side).
Start of impregnation
End of impregnation
Fig. 5.16. Flow through an array of channels. Reference geometry. Lower side (preform side).
Inlet
Outlet
Outlet
Inlet
- 125 -
Results show that voids are created underneath every cell from the start of the
injection until the end, although the velocity of the flow decreases during the
injection.
Table 5.9. Boundary conditions. Flow through an array of diamond cells.
Parameter Value
Cell shape diamond 90
Cell size 2 mm side length
Channel width 1 mm
Thickness 1 mm
Bag dimensions 12.69 mm * 12.69 mm * 1 mm
Bag Porosity 0.33
Bag permeability 1*10
-8
m
2
P
inlet
1*10
5
Pa
P
outlet
0 Pa
Reinforcement thickness 3 mm
Reinforcement permeability 1*10
-12
m
2
Reinforcement porosity 0.6
5.5. Proposed plan for future work
5.5.1. Modelling
In the preliminary study described earlier in this chapter, CFD analysis was used to
model flow through a unit cell. In addition examples of flow through an array of
channels and cells were given. This section proposes to extend the study to the
modelling of flow through a complete membrane. In addition a study to model best
infusion strategy is proposed.
Modelling flow in a vacuum membrane
This phase of the study involves modelling the infusion of flat geometries where a
network of cells and channels represents the new vacuum membrane. In this case
more than one unit cell is modelled. The size of the cells and channels is varied and
two types of analysis are conducted. One considers one size of cells and channels at
a time and the other varies the size of the cells and channels in different areas of
the bag, thus providing a range of permeabilities. Therefore it is possible to analyse
flow behaviour through a distribution medium of non-uniform permeability and
study its effect on lead-lag and flow front profile. A preform is modelled underneath
the bag and the study is conducted for a range of preform permeabilities. Another
test consists of adjusting the permeability of the preform to that of the bag so a
constant permeability ratio is kept between the bag and the preform. The aim is to
- 126 -
investigate the occurence of a uniform lead-lag and flow front profile throughout the
preform. This study can be extended to the modelling of non-flat geometries where
the permeability of both the new distribution medium and the preform are varied
according to the geometry of the part.
Modelling infusion strategy
The other interest in designing such a bag is the possibility to simplify infusion
strategies. Usually the infusion of a preform is modelled using different infusion
strategies in order to predict the minimum infusion time and avoid the formation of
dry patches. Depending on the geometry of the part, the number of infusion points
can vary in order to prevent predicted dry patches from occuring. With the new
membrane design, the eventual occurrence of dry patches is reduced and therefore
infusion strategies can be simplified and the number of infusion inlets reduced. A
study involves the modelling the infusion of a preform with either a classic
distribution medium on top of the preform or the new membrane geometry. Then
different infusion strategies are used in both cases and the best strategy is
determined for each configuration. The best strategies for both cases are compared.
The interest is to demonstrate that the infusion strategy would be simpler with the
new membrane design and the occurrence of dry patches reduced. Figure 5.19
shows a summary of the modelling study (preliminary work and proposed work).
5.5.2. Experimental analysis
This section describes experiments that could be conducted to validate the
modelling work. The vacuum infusion process is used to conduct the experiments.
Infusion under a rigid cover
Flow through a mould cavity made of two rigid halves is analysed. The upper cover
is made of undeformable glass. Polyamide diamond cells of selected dimensions and
thickness are adhered to the surface. It is important to notice that the space
between the cells determines the size of the channels. Therefore they need to be
accurately positioned on the glass plate. The base of the mould is an undeformable
glass plate. The mould cavity is sealed using a vacuum bag and air is extracted via
a pump. Two different types of experiments are conducted: one consists of injecting
- 127 -
fluid through an array of channels only,in the other experiment, the fluid is injected
through both channels and a preform located underneath the cells. The interest of
the first series of experiment is to compare results for flow modelling through a unit
cell to experimental flow through a network of channels. Fill time and resin
consumption are the parameters of interest to validate the study.
Fig. 5.19. Diagram: modelling strategy for new membrane using CFD.
The second type of experiments investigates void creation in the reinforcement
underneath the cells. According to reinforcement permeability and architecture
conclusions can be made about the influence of dimensional and structural
parameters on void formation. Different sizes of cell and channel space are tested
as well as different reinforcement architectures in order to compare results with CFD
analysis.
Unit cell Membrane
Vary preform
permeability
Infusion time / Void formation / Lead-lag
Preform permeability
constant
Vary cell size
Vary channel size
Reference geometry
No preform
1 cell size
1 channel size
Infusion time identical
to distribution
medium
Model infusion strategies
Infusion
time in
distribution
medium
Infusion
time in new
membrane
Validate efficiency
of new membrane
- 128 -
In addition a flow monitoring system, like SmartWeave, can be used to record
flow progression through the thickness of the preform in order to study void size
during impregnation. In this case the fluid injected has to be conductive in order to
monitor flow progression. A description of the SmartWeave system is available in
chapter 2 section 2.3.
Infusion under a soft membrane
In the next step the same experiments presented above are conducted but the glass
cover is replaced by a soft deformable membrane. Such a membrane can be
produced in-house using a hand-made mould. The membrane can be moulded in
any rubber material such as silicone or latex. The cells are directly moulded with the
membrane. As proposed previously, the membrane is used to conduct two types of
experiments. One consists of injecting a fluid through the channels of the
membrane only. The other involves the impregnation of a preform underneath the
membrane. Again a flow monitoring system is used to record flow progression
through the thickness of the preform and indicate the presence of voids.
It is suggested that the experiments described above are conducted for different
types of reinforcement and different preform thicknesses so that the analysis is
validated for a range of fibre architectures and permeabilities.
After this first series of experiments, a domain of solution for bag geometry
according to preform permeability is obtained. Then the study focuses on flow
through a membrane geometry of varying permeability. It is suggested to mix
different sizes of channels and cells which are all part of the domain of solution
previously determined, both analytically and experimentally. The experiments are
conducted first on a flat plate and for different types of reinforcement architecture
and permeability. The next step involves the infusion of preforms in moulds of
varying geometries so that permeability changes also occur in the preform. Again
the experiments need to be conducted for different types of reinforcement
architectures. Figure 5.20 shows a summary of the proposed experimental analysis.
- 129 -
5.5.3. Extension to general applications
The determination of the best geometry for a bag can be extended to any general
application. When designing a vacuum bag, the only information available is likely to
be the average thickness of the dry reinforcement, the permeability of the
reinforcement and the thickness of the membrane. With this information, the first
element that can be determined is the ideal membrane permeability required.
Several solutions will then be possible and the best one will be for a size of cells and
channels which prevents or minimises voids creation. The next step will consist of
determining the best cell size according to the reinforcement thickness and the
reinforcement permeability. Again some solutions will be eliminated. The best
solution will then be the one for which the lead-lag is minimal. Figure 5.21 describes
the different steps of the study proposed.
Fig. 5.20. Diagram: experimental analysis of new membrane.
Rigid cover Flexible membrane
No preform Preform
1 cell size
1 channel size
Several cell sizes
Several channel sizes
Fill time
Resin consumption
Void formation
Lead-lag
Vary reinforcement
architecture
Vary preform
permeability
1 cell size
1 channel size
Several cell sizes
Several channel sizes
Fill time
Resin consumption
- 130 -
Fig. 5.21. Diagram: general methodology to design a vacuum membrane.
5.6. Conclusions
This chapter presents a design methodology to develop an innovative vacuum
membrane. The study shows the interest in developing a new type of vacuum bag
where the flow enhancement properties of the distribution medium are included in
the vacuum membrane. Results from a preliminary analysis show that it is possible
to model and predict the effect of preform permeability and thickness compared to
that of the membrane geometry on infusion time and void creation.
Preform
thickness
Bag thickness Preform
permeability
Vary cell size
Bag
permeability
No voids
Voids
Several
solutions
Vary cell size
Minimum
lead-lag
One solution
Known
information
- 131 -
The next step of the study would involve a series of experiments where a preform is
impregnated underneath the cells. Then conclusions could be made on void creation
and results could be compared to flow predictions. The experimental analysis could
move from the use of a solid glass plate and plastic cells to a soft membrane with
soft cells embedded on its surface. In addition the size of cells and channels could
vary on the surface of the mould in order to suit the permeability of the preform
underneath. Conclusions could be made about the efficiency of such a design and
its advantages for flow prediction, infusion strategy set-up, flow control, material
and production cost (see Appendix 9) as well as process repeatability and quality
consistency.
- 132 -
References
1. C. Williams, J. Summerscales, S. Grove.Resin infusion under flexible tooling
(RIFT): a review. Composites Part A, 27, 1996, p. 517-524.
2. D. Heider, A. Karamitsos, J. W. Gillespie, S. Walsh, E. Rigas, W. Spurgeon, W.
Roy. Experimental validation and optimisation of the FASTRAC process. 22
th
International SAMPE Europe conference, April 2001, Paris, France, p. 293-303.
3. R.S. Parnas, A.J. Salem, T.A.K. Sadiq, H-P Wang, S.G Advani. The interaction
between micro- and macroscopic flow in RTM preforms. Composite structures,
27, 1994, p. 93-107.
4. T.A.K. Sadiq, S.G. Advani, R.S. Parnas. Experimental investigation of
transverse flow through aligned cylinders. International Journal Multiphase
Flow, 21, N
o
5, 1995, p. 755-774.
5. R.S. Parnas, F.R. Phelan. The effect of heterogeneous porous media on mold
filling in resin trasnfer molding. SAMPE Quarterly, Jan 1991, 22, p. 53-60.
6. M.A.A. Spaid, F.R. Phelan Jr. Modelling void formation dynamics in fibrous
porous media with the lattice Boltzmann method. Composites Part A, 29,
1998, p. 479-755.
7. D. Abraham, R. Mclhagger. Investigations into various methods of liquid
injection to achieve mouldings with minimum void contents and full wet out.
Composites Part A, 29, 1998, p. 533-539.
8. N. Pearce, F. Guild, J. Summerscales. A study of the effects of convergent flow
fronts on the properties of fibre reinforced composites produced by RTM.
Composites Part A, 29, 1998, p. 141-152.
- 133 -
Chapter 6
Conclusions and future work
6.1. Conclusions
This experimental characterisation of vacuum infusion allowed a better
understanding of the process and material parameters involved in preform infusion
as well as to show possibilities for process improvements.
Study of process parameters shows that:
- Bagging materials strongly influence the infusion of the preform in term of
compaction and time. Some bagging strategies result in masking the flow
enhancement or compaction properties of each material taken individually.
- All types of vacuum bags are not adequately designed for being used in VI
(thickness / elongation, surface pattern).
The analysis of the material parameters shows that:
- The porosity permeability relationship as well as the compliance of both the
preform and the distribution medium must be compared prior to choosing the
best suited distribution medium which provides the required level of flow
enhancement in the preform.
- Both the compressibility relaxation and the evolution of compaction with time
under a given pressure must be charaterised for infused preforms in order to
predict the use of bleeding technique to end the infusion. However, the study
presented in this report shows that the latter is not a prerequisite.
In addition, analysis of flow in VI, when using a distribution medium, reveals that
three major phases occur during preform impregnation and that both in-plane and
trans-plane permeability of the preform as well as in-plane permeability of the
distribution medium are involved in the impregnation process:
- 134 -
- The first phase occurs at the onset of the infusion where in-plane flow in both
the distribution medium and the preform lead the impregnation. Trans-plane
flow has already occurred in the preform and a lead-lag is created between the
distribution medium and the preform.
- The second phase occurs when trans-plane flow reaches the transversal
extremity of the preform. Lead-lag reaches its maximum and trans-plane flow
leads the impregnation at the front of the preform while flow is still in-plane in
the distribution medium. Lead-lag decreases due to a shorter trans-plane
distance to cover through the preform compared to in-plane dimensions.
- The third phase corresponds to the occurrence of a steady-state lead-lag. The
latter is linked to the velocity of the flow which depends on both preform and
distribution medium permeability.
Study of flow through a HDPE stitched-chain distribution medium reveals that:
- This type of material can be modelled as an open channel.
- Both nesting and bag bending affect significantly its permeability in opposite
ways. Nesting results in a decrease of the porosity of the distribution medium
and the effect increases with increasing the number of layers. Bag bending
decreases the overall porosity of the infused cavity too but the effect decreases
when increasing the number of material layers.
This leads to the possibility of designing an integrated vacuum membrane where
arrays of channels are created between embedded cells, thus creating a distribution
medium at the surface of the membrane. Modelling of flow through such a structure
reveals that void creation in the preform can be predicted. The permeability and
thickness of the preform related to the permeability of the channels in the
membrane are seen to be the key factors in controlling the occurrence of voids in
the preform during infusion. Modelling analysis shows that the occurrence of voids
in a preform can be avoided. However, when voids occur flow modelling analysis
cannot predict the evolution of the location and size of the voids.
- 135 -
6.2. Future work
Following this experimental analysis of VI further analysis and development of
several aspects of the process can be envisaged:
- Bleeding: it would be useful to develop a model for bleeding that takes into
account for preform compaction and time as well as for the mass of resin flowing
in and out of the mould cavity so that the best strategy to end an infusion could
be predicted.
- Permeability measurement: knowledge of the quantity of resin flowing into the
preform during infusion could be used to measure the permeability of the fibre
stack on-line during VI. Indeed the mass of resin in an impregnated section of
the part could be related to the porosity of the preform, assuming that no voids
occur and that all the space available between the fibres is occupied by the
resin. Knowing the pressure gradient applied on the length of this preform
section would then be enough to determine the permeability of that section of
the part using Darcys law. The permeability calculations would be more or less
accurate according to the length of the section analysed and results could be
compared to in-plane permeability values from measurements made in fixed
thickness cavity with imposed flow velocity.
- Lead-lag: it would be useful to know the direction of the velocity vectors during
the impregnation and especially at the transition between in-plane flow behind
the flow front and trans-plane flow at the front. This information would help in
determining the real distance covered by the resin before it reaches the lower
side of the preform. A start for such a study could be achieved in a simple way
by comparing the thickness of the preform at the front with the thickness of a
preform under a similar compaction level. The thickness of the preform being
impregnated would be determined using a form of Darcys law: assuming that
the porosity and permeability of the preform are known the distance covered by
the fluid between the upper and lower side of the preform is determined using
recorded flow front positions with time. This distance which corresponds to the
thickness of the preform is then compared to the thickness of a preform of same
compressibility / relaxation level. If both thicknesses are identical then veloctiy
vectors are exactly vertical to the preform. However, if the thicknesses are
- 136 -
different it means that the velocity vectors form an angle with the vertical of the
preform. Although this method would allow us to know whether the velocity
vectors at the front are vertical to the preform or not it would not give any
information on the real inclination of the vectors. In addition this information is
not sufficient since even if the velocity vectors are vertical at the front the
transition between in-plane flow and trans-plane flow is not straightforward and
the evolution of velocity vectors with flow position is not given. Knowing this
would give information of the shape of the lead-lag during infusion.
- Novel vacuum membrane: further analysis is required in order to validate the
efficiency of such a membrane, as presented in this report. In addition a
systematic analysis needs to be developed which would allow designing a
specific vacuum bag to every part infused as a function of the permeability of
the preform and the geometry of the part.
- Distribution medium: The study of the distribution medium could be conducted
further by analysing flow phenomena in the preform when the distribution
medium is located at places different from the top of the perform (between
layers of reinforcement for example). Properties of the distribution medium
should also be compared to the flow enhancement properties of other
reinforcements of equivalent permeability. This would also give more information
on the lead-lag and results from this study could be analysed further (lead-lag
creation, effect to in-plane and trans-plane flow).
- 137 -
Appendices
- 138 -
Appendix 1
Permeability measurements
1. In-plane permeability
1.1. Method
The in-plane permeability of the different reinforcements was measured using a
radial flow experiment. The procedure is an evolution of that outlined by Chick [1]
and developed by Duarte [2]. The analysis is based on Darcys law with constant
flow rate. Assuming that the axes of measurements corresponds to the principal
axes permeabilities then the following relations can be deduced for the wetting
(transient) and wetted (steady-state) permeability values:
Steady-state permeability:
|
|
.
|

\
|


=
x
x
xx m
x x
r
r
K h
Q
P P
0
0
ln
2 t
| u
1
|
|
.
|

\
|


=
y
y
yy m
y y
r
r
K h
Q
P P
0
0
ln
2 t
| u
2
Transient permeability: 3
4
Where P is the pressure, Q the flow rate, r
o
the radius of the injection point, r
x
and
r
y
the radius of the flow front on x and y axes respectively at any time t, h is the
|
|
|
|
.
|

\
|
+



=
2
0
ln
2
r
h
t Q
r
K h
Q
P
x
xx
x
|
t c
| t
| u
|
|
|
|
.
|

\
|
+



=
2
0
ln
2
r
h
t Q
r
K h
Q
P
y
yy
y
|
t c
| t
| u
- 139 -
thickness of the stack and K is the permeability of the material on x (K
xx
) and y
(K
yy
) axes.
| is the flow front aspect ratio:
y
x
r
r
= | 5
Since the pressure transducers are located along the principal axes in this work, |
can be calculated using the coordinates of the pressure transducers.
1.2. Apparatus
The apparatus, shown in Figure 1. a), consisted of a 23 mm aluminium plate with a
400 mm diameter, 2.1 mm high cavity. A 10 mm diameter central injection port and
2 peripheral vents controlled fluid flow. An aluminium cover closed the cavity and a
bolted supporting frame restrained deflections. Four pressure transducers within the
base plate and 1 in the cover were monitored continuously, whilst mineral oil, HDX
30 from Trent Oils, was injected at constant flow rate using a servo-hydraulic
testing machine as the actuator. Fabric samples were cut to size, stacked to
maintain orientation and a 10 mm hole was punched in the assembly to minimise
entry losses. Experiments were run using actuator speed of 10 to 20 mm/min,
corresponding to flow rates of 3.35*10
-6
to 6.70*10
-6
m
3
/s. The thickness of the
cavity was adjusted using peripheral shims. The viscosity of the oil was measured
using a viscometer Brookfield LV-DII. The temperature of the oil in the feeding tank
was also recorded and results were fitted to the viscosity versus temperature curve
obtained from recordings on the viscometer. Figure 2 shows that temperature was
fairly constant throughout the test, between 20 and 21C, as the testing room was
air conditionned. Consequently the maximum variation in viscosity was only 10
mPa.s. The curve does not show as many points as the number of experiments as
many of the recordings shown the same viscosity - temperature value.
1.3. Test matrix
The in-plane permeability of the three reinforcements involved in the experimental
analysis was measured. For each reinforcement the in-plane permeability was
measured for three different fibre volume fractions as shown in Table 1. The fibre
- 140 -
volume fraction V
f
was determined with knowledge of the fibre glass density,
number of fabric layers and cavity thickness as shown in Equation 6.
f
de
f
h
n S
V
p

= 6
a). Apparatus.
b). Location of pressure transducers and injection port.
Fig.1. In-plane permeability measurements: apparatus
Circular cavity
located 2.1 mm
below peripheral
plate
Cover
Injection port
Pressure
transducer
Injection port
73575
57833
52128
67754
56656
- 141 -
Fig. 2. Viscosity of the oil HDX 30 measured during in-plane permeability testing.
Table 1. Test matrix: in-plane permeability measurements
Reinforcement
Number of
layers
Cavity thickness (mm) V
f
(%)
Actuator speed
(mm/min)
U 751 / 375
5
6
7
3.1
3.1
3.1
20.0
28.5
30.0
20
20
20
RT 600
3
4
5
2.6
2.6
2.6
20.0
30.0
40.0
20
20
10
FGE 117
2
2
3
2.6
3.1
2.6
29.0
35.0
44.0
20
20
10
1.4. Pressure transducers calibration
Prior to starting permeability measurements the pressure transducers were
calibrated in order to ensure accurate recordings. Figure 1. b) shows the
identification number of each transducer and its location on the permeability rig.
The test consisted of recording readings of each transducer according to different
pressure loadings. Figure 3 displays the resulting pressure as a function of voltage
reading. The equation of each curve was then used to determine the corresponding
pressure values P from the voltage U recorded during permeability experiments.
256
258
260
262
264
266
268
270
19.8 20 20.2 20.4 20.6 20.8 21 21.2
Temperature (C)
V
i
s
c
o
s
i
t
y
(
m
P
a
.
s
)
- 142 -
Fig. 3. Evolution of pressure transducer readings
as a function of the applied pressure.
1.5. Results
Figures 4 - 6 show both steady-state and transient in-plane permeability of each
reinforcement as a function of fibre volume fraction. As expected the continuous
filament random mat offers the highest permeability, while the triaxial fabric shows
the lowest. Steady-state permeability is higher than transient for all fabrics
considered. This highlights the importance of fabric lubrication and its effect on
preform permeability changes during VI process.
Fig. 4. In-plane permeability of triaxial fabric FGE 117.
0
20
40
60
80
100
120
140
0 2000 4000 6000 8000 10000 12000 14000
Volts
P
S
I
Transducer 67754
Transducer 57833
Transducer 56656
Transducer 73575
Transducer 52128
P=aU + b
Transducer reading (Volts)
P
r
e
s
s
u
r
e
(
P
S
I
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0.32 0.44 0.48
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0.32 0.44 0.48
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0.32 0.44 0.48
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
0.32 0.44 0.48
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
Transient Kx
Transient Ky
Steady-state Kx
Steady-state Ky
- 143 -
Table 2. Comparison of steady-state and transient permeability
Reinforcement K
Steady-state
/ K
transient
U 751 / 375 3
RT 600 2
FGE 117 - Parallel orientation to 0 axis 2.15
FGE 117 - Perpendicular orientation to 0 axis 0.75
Fig. 5. In-plane permeability of continuous filament mat U 751 / 375.
Fig. 6. In-plane permeability of plain weave RT600.
0
10
20
30
40
50
60
70
80
90
0.32 0.4 0.491
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0
10
20
30
40
50
60
70
80
90
0.32 0.4 0.491
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0
10
20
30
40
50
60
70
80
90
0.32 0.4 0.491
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0
10
20
30
40
50
60
70
80
90
0.32 0.4 0.491
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
Transient Kx
Transient Ky
Steady-state Kx
Steady-state Ky
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
0.27 0.36 0.45
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
0.27 0.36 0.45
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
0.27 0.36 0.45
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
0.27 0.36 0.45
Fibre volume fraction
K
(
1
0
-
1
0
m
2
)
Transient Kx
Transient Ky
Steady-state Kx
Steady-state Ky
- 144 -
2. Trans plane permeability
2.1. Method
The steady-state trans-plane permeability of the different reinforcements was
measured using a one-dimensional rectilinear flow experiment. The analysis is
based on Darcys law with constant flow rate as described in Equation 7 [3-5]. The
method consists of recording the pressure difference created by the resistance of
the fluid to flow through the thickness of the fabric. Pressure recordings according
to the position of the pressure transducers, velocity of the fluid and porosity of the
sample are used to deduce the trans-plane permeability value.
P S
e Q
K
zz


=
u
7
2.2. Apparatus
The apparatus, shown in Figure 7 consisted of two 250 mm (external diameter)
aluminium cylinders which formed the upper and lower parts of the tool. A 80 mm
diameter cavity, respectively 25 mm and 75 mm deep, admitted the fluid in each
cylinder. Two pressure transducers, located on the upper and lower cylinder
respectively, were monitored continuously, whilst mineral oil, HDX 30 from Trent
Oils, was injected at constant flow rate using an Instron 1195 servo-hydraulic
testing machine as the actuator. A sketch of the set-up in shown in Figure 8. An
injection port and a vent, located in the upper and lower cylinder respectively,
controlled fluid flow. The lower cavity admitted a 80 mm cylinder (50 mm high, 35
mm thick) which surrounded the upper cavity. 10 to 30-layer samples of 100 mm
diameter were placed in the lower cavity and preform thickness was adjusted by
measuring the distance separating the two main cylinders. Two aluminium grids,
surrounding the test preform, prevented its deflection under fluid pressure. O-rings
were located in the lower cavity and outside upper cylinder to prevent leakage.
The two cylinders were assembled using four M16 screws. Experiments were run
using actuator speed from 10 to 100 mm.min
-1
, corresponding to flow rates of
3.35*10
-6
to 33.50*10
-6
m
3
.s
-1
. Voltages from the two pressure transducers were
recorded through a Visual Designer program. Measurements were stopped when
the fluid reached the vent and voltages readings were constant.
- 145 -
Fig. 7. Trans-plane permeability measurement apparatus.
2.3. Test matrix
Trans-plane permeability of the three reinforcements involved in the experimental
analysis was measured. Table 3 shows the tests conducted.
2.4. Pressure transducers calibration
As for in-plane permeability measurements, the pressure transducers were
calibrated prior to experiments. Table 4 shows the calibrating values of the two
transducers involved in the measurements.
Cavity
thickness
Pressure
transducer P
2
Inlet
Outlet
Pressure
transducer P
1
Cavity
Fabric
sample
Lower cylinder
Grid
Upper
cylinder
Cylinder
Grid
- 146 -
Fig. 8. Trans-plane permeability measurement set-up.
Table. 3. Test matrix: trans-plane permeability measurement.
Reinforcement
V
f
(%)
Cavity thickness
(m)
Actuator
speed
(mm/min)
Flow rate
(m
3
/s)
Number of
layers
25 0.01417 100 33.50*10
-6
24
30 0.01181 100 33.50*10
-6
24
35 0.01012 100 33.50*10
-6
24
40 0.00886 100 33.50*10
-6
24
45 0.00787 100 33.50*10
-6
24
50 0.00709 100 33.50*10
-6
24
U 751 / 375
55 0.00644 100 33.50*10
-6
24
40 0.01771 50 16.75*10
-6
30
50 0.01417 50 16.75*10
-6
30 RT 600
60
0.01181
20 6.70*10
-6
30
25 0.00600 100 33.50*10
-6
10
30 0.00500 100 33.50*10
-6
10
35 0.00600 50 16.75*10
-6
14
45 0.00600 20 6.70*10
-6
15
FGE 117
50 0.00500 10 3.35*10
-6
15
Table 4. Pressure transducers calibration
mV to PSI P = aU + b
Tranducer a b
P
1
0.0208 2.5135
P
2
0.0127 0.2301
PSI to Pa 1PSI = 6901.2 Pa
Instron machine
Bucket
Apparatus
Sample
Piston
- 147 -
0.0
0.5
1.0
1.5
2.0
2.5
0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60
Fibre volume fraction
K
(
1
0
-
1
1
m
2
)
2.5. Results
Figures 9 to 11 show the trans-plane permeability of the different reinforcements
tested. As expected the CFRM shows the higher trans-plane permeability while the
triaxial fabric and plain weave show lower values.
Fig. 9. Steady-state trans-plane permeability of
continuous filament random mat U 751 /375.
Fig. 10. Steady-state trans-plane permeability of plain weave glass fabric RT 600.
0.0
10.0
20.0
30.0
40.0
50.0
60.0
70.0
80.0
90.0
0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60
Fibre volume fraction
K
(
1
0
-
1
1
m
2
)
- 148 -
Fig. 11. Steady-state trans-plane permeability of triaxial glass fabric FGE 117.
This result is expected and is due to the architecture of each fabric. The CFRM glass
is much more porous than the two other reinforcements while the triaxial fabric and
the plain weave are woven fabrics with a higher surface density. In addition Table 5
shows a comparison of in-plane and trans-plane permeability according to whether
the in-plane permeability is steady-state or transient. Results show that the
differences between in-plane and trans-plane permeabilities vary significantly
according to the type of reinforcement. Highest variations occur for the triaxial
fabric, for flow parallel to the 0 fibres, followed by the CFRM. Differences in
permeability are much higher between steady-state and trans-plane permeability
than transient and trans-plane permeability for all reinforcements. This is expected
since in-plane steady-state permeability is generally higher than transient
permeability due to the lubrication of the fabric during impregnation.
Table 5. Comparison of steady-state and
transient in-plane permeability with steady-state trans-plane permeability
Reinforcement K
Steady-state
/ K
Trans-plane
K
Transient
/ K
Trans-plane
U 751 / 375 26 9
RT 600 5.5 2.75
FGE 117 - Parallel orientation to 0 tow 32 14
FGE 117 - Perpendicular orientation to 0 tow 11 8
0.0
0.5
1.0
1.5
2.0
2.5
0.20 0.25 0.30 0.35 0.40 0.45 0.50 0.55 0.60
Fibre volume fraction
K
(
1
0
-
1
1
m
2
)
- 149 -
References
1. J.P. Chick, C.D. Rudd, P.A. van Leeuwen, T.I. Frenay. Material Characterization
for Flow Modeling in Structural Reaction Injection Molding. Polymer Composites,
February 1996, 17, N
o
1, p. 124-136.
2. A. Duarte. In-plane permeability of interleaved preforms for resin transfer
moulding.
3. C-H. Wu, T. J. Wang, L.J. Lee. Trans-plane fluid permeability measruement and
its applications in liquid composite molding. Polymer Composites, August 1994,
15, N
o
4, p. 289-298.
4. T.D. Papathanasiou, P.D. Lee. Morphological effects on the transverse
permeability of arrays of aligned fibers. Polymer Composites, April 1997, 18, N
o
2, p. 242-253.
5. I. C. Visconti, A. Langella, M. Durante. Analysis of transversal permeability for
different types of glass fiber reinforcement. Applied Composite Materials, March
2003, 10, N
o
2, p. 119-127.
- 150 -
Appendix 2
Materials characteristics
1. Reinforcements
Table 1. Technical characteristics: E-glass reinforcements
Description Triaxial fabric Plain weave
Continuous
filament
random mat
Material E-Glass E-Glass E-Glass E-Glass
Commercial name FGE 117 E-TLX 1169 RT 600 U 751 / 375
Supplier Formax Cotech Vetrotex Vetrotex
Total weight (g.m
-2
) 1173 1169 600 375
Ply construction 0/-45/+45 0/-45/+45 0 / 90 Random
Weight / axis (g.m
-2
) 573 / 300 / 300 567 / 301 / 301 300 / 300 N / A
% by weight 48 / 26 / 26 48 / 26 / 26 50 / 50 N / A
2. Fluids
Table 2. Technical characteristics: fluids
Description Oil Polyester resin Initiator
Product name HDX 30 701 PAX M50
Supplier Trent oils Scott Bader Akzo Chemie
Material Mineral oil Isophthalic polyester
Methyl ethyl ketone peroxide
35% in phthalate plasticizer
Viscosity 190 mPa.s (25C) 160 mPa.s (25C)
20 mm
2
/s (20C)
(Kinematic viscosity)
- 151 -
3. Consumables
Table 3. Technical characteristics: consumables
Description Vacuum bag Bleed out fabric
Distribution
medium
Product name Embossed B205 60B H8036
Supplier Tygavac Tygavac Tygavac Palmhive Textiles
Material Polyamide Polyamide
Woven
polyamide
HDPE
Material density (g.cm
-3
) 0.945
Total weight (g.m
-2
) 100
Filament diameter (mm) 0.25
Maximum temperature
resistance (
o
C)
205 205 205 80
Tensile strength (MPa) 85-105 45-65 N/A N/A
Maximum elongation (%)
200
400 N/A N/A
Thickness (mm) 0.3 0.05 0.1 ~ 0.6
1
1
The thickness of the distribution medium was measured from a sample of the material cast into resin. A
section of the cast was cut and polished and the thickness of the distribution medium was measured
under an optical microscope at magnification G*50. An average value from 20 measurements was taken
to determine the average thickness of the distribution medium.
- 152 -
Appendix 3
Viscosity of fluids
1. Mineral oil
Fig. 2. Viscosity of mineral oil HDX30 during Tests 1 to 8, Chapter 2.
2. Polyester resin
Fig. 4. Viscosity of catalysed 701 PAX polyester resin as a function of catalyst loading.
0
200
400
600
800
1000
1200
1400
0 10 20 30 40 50 60
Time (min)
V
i
s
c
o
s
i
t
y
(
m
P
a
.
s
)
1% b.w
0.75% b.w
1.25% b.w
0
50
100
150
200
250
300
19 20 21 22 23 24 25
Temperature (
o
C)
V
i
s
c
o
s
i
t
y
(
m
P
a
.
s
)
Test 1
Test 2
Test 3
Test 4
Test 5
Test 6
Test 7
Test 8
- 153 -
Fig. 5. Temperature of catalysed 701 PAX polyester resin
as a function of catalyst loading.
21
22
23
24
25
26
27
28
29
0 10 20 30 40 50 60
Time (min)
T
e
m
p
e
r
a
t
u
r
e
(
C
)
1% b.w
0.75% b.w
1.25% b.w
- 154 -
Appendix 4
Thickness distribution in laminates at the end of infusion
The thickness is represented by contour levels, in a range of colours going from blue
for low thickness to red for high thickness. Results are presented as follows: the
contour levels of each individual reinforcement were set to the same minimum and
maximum value so that results could be compared, as described in Table 1. Results
are presented in three-dimensional curves where the thickness of the laminate is
displayed on the Z axis, the distance from the inlet is set on the X axis and the Y
axis represents the width of the laminate.
Table 1. Thickness contour levels specific to each fabric type.
Test Reinforcement Number of
contour
levels
Lower thickness
contour level (mm)
Higher thickness contour
level (mm)
9 U 751 / 375 40
10 U 751 / 375 40
11 U 751 / 375 40
4 7
12 FGE 117 40
13 FGE 117 40
14 FGE 117
40
6 8
15 RT 600
40
16 RT 600
40
17 RT 600
40
3 5
- 155 -
Test 9 Inlet closed for 20 minutes
Test 10 Inlet closed for 10 minutes
Test 11 Inlet closed for 2 minutes
Fig. 1. Thickness distribution in U751/375. Tests 9 to 11.
- 156 -
Test 12 Inlet closed for 10 minutes
Test 13 Inlet closed for 2 minutes
Test 14 Inlet left open until resin gelation
Fig. 2. Thickness distribution in FGE 117. Tests 12 to 14.
- 157 -
Test 15 Inlet closed for 10 minutes
Test 16 Inlet closed for 2 minutes
Test 17 Inlet left open until resin gelation
Fig. 3. Thickness distribution in RT 600. Tests 15 to 17.
- 158 -
U 751 / 375 - Test 10
FGE 117 - Test 12
RT 600 - Test 15
Note: Contour levels were set to identical minimum (3 mm) and maximum (8 mm).
Fig. 4. Comparison of thickness distribution between the different types of reinforcement.
Inlet closed 10 min before resin gelation.
- 159 -
Appendix 5
Effect of outlet pressure on lead-lag
The effect of outlet pressure on lead-lag was measured by conducting 3 series of
experiments. 9-layer preforms of E-TLX 1169 triaxial fabric were infused at 3
different absolute pressures: 101.3 mbar, 401.3 mbar and 701.3 mbar, i.e vacuum
of respectively 900 mbar, 600 mbar and 300 mbar in the cavity. The preform
measured 700 mm by 300 mm with linear flow along the 700 mm side (see Table
1).
Table 1. Test matrix: effect of outlet pressure on lead-lag
Fabric Distribution medium
Layers Orientation Layers Stitch orientation
Outlet absolute
pressure (mbar)
9 0 parallel 1 Weftwise 101.3
9 0 parallel 1 Weftwise 401.3
9 0 parallel 1 Weftwise 701.3
This series of experiments confirmed that trans-plane permeability is not influenced
by the pressure gradient. These consisted in infusing a 9-layer triaxial preform
under respectively 101.3 mbar, 401.3 mbar and 701.3 mbar of absolute pressure.
Figure 1 shows the evolution of infusion time with flow front distance for the three
tests. Figures 2 and 3 show the evolution of flow velocity for the upper and lower
faces of preforms infused under the three different absolute pressures. The upper
velocities are higher than the lower velocities for all pressure levels. Velocities on
both the upper and lower faces are higher at 101.3 mbar of absolute pressure and
decrease when absolute pressure increases.
However the effect of pressure on velocity is more significant for flow on the upper
face of the preform than flow on the lower face. This demonstrates that the
pressure applied in the cavity affects in-plane permeability and confirms that trans-
plane permeability is not affected by pressure gradient.
x
llmax
obtained at the beginning of the infusion occurs at the same distance from the
inlet and is identical whatever the absolute pressure (see Figure 4). This result
shows that the triaxial fabric has a low compaction level and therefore preform
thickness does not vary significantly with pressure level. However the time required
to reach x
llmax
is shorter under 101.3 mbar and increases with absolute pressure.
- 160 -
Fig. 1. Time as a function of flow front position on both upper and lower faces of the
preform, for different absolute pressures. Trendlines are polynomial curve fit.
The absolute pressure affects more infusion time than lead-lag for a given preform thickness.
Fig. 2. Fluid velocity as a function of flow front position
on both upper and lower faces of the preform for different vacuumlevels.
Flow velocity is much higher on upper side of the preform than lower side. In addition it is
generally higher for thinner preforms. However the velocity decreases quickly on the upper side
with flow front progression and after 200 mm all velocities become closer to eachother.
0
500
1000
1500
2000
2500
0 100 200 300 400 500 600
Flow front position (mm)
T
i
m
e
(
s
e
c
)
701.3 mb upper
701.3 mb lower
401.3 mb upper
401.3 mb lower
101.3 mb upper
101.3 mb lower
0
2
4
6
8
10
12
14
16
18
0 100 200 300 400 500 600
Flow front position (mm)
V
e
l
o
c
i
t
y
(
m
m
/
s
e
c
)
701.3 mbar upper
701.3 mbar lower
401.3 mbar upper
401.3 mbar lower
101.3 mbar upper
101.3 mbar lower
T
i
m
e
(
s
)
V
e
l
o
c
i
t
y
(
m
m
/
s
)
- 161 -
Fig. 3. Fluid velocity as a function of flow front position on both upper and lower faces of the
preform between 300 mm and 500 mm from the inlet and for different vacuum levels.
Trendlines are polynomial curve fit.
This is due to the fact that flow rate is a direct function of the vacuum level,
assuming that the permeability of the preform is constant. In the present case
vacuum level affects the in-plane permeability more than it affects trans-plane
permeability. Consequently vacuum level affects in-plane flow rate in the
distribution medium, which therefore has an effect on the trans-plane flow rate in
the preform.
Fig. 4. Lead-lag expressed as a difference in distance and
for different vacuum levels. Trendlines are polynomial curve fit.
The absolute pressure does not affect lead-lag.
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
250 300 350 400 450 500 550
Flow front position (mm)
V
e
l
o
c
i
t
y
(
m
m
/
s
e
c
)
701.3 mbar upper
701.3 mbar lower
401.3 mbar upper
401.3 mbar lower
101.3 mbar upper
101.3 mbar lower
0
20
40
60
80
100
120
140
160
180
0 100 200 300 400 500 600
Flow front position (mm)
L
e
a
d
-
l
a
g
(
m
m
)
701.3 mbar
401.3 mbar
101.3 mbar
V
e
l
o
c
i
t
y
(
m
m
/
s
)
- 162 -
Appendix 6
Effect of varying preform thickness on lead-lag
Fig. 1. Time as a function of flow front position on both upper and lower faces of the preform.
Comparison between preforms of constant thickness and preforms of increasing thickness.
Trendlines are polynomial curve fit.
Fig. 2. Time as a function of flow front position on both upper and lower faces of the preform.
Comparison between preforms of constant thickness and preforms of increasing thickness.
Trendlines are polynomial curve fit.
0
200
400
600
800
1000
1200
1400
1600
1800
0 100 200 300 400 500 600 700
Flowf ront position (mm)
T
i
m
e
(
s
e
c
)
8/11 layers upper
8/11 layers lower
Polynomial (11 layers lower)
Polynomial (11 layers upper)
Polynomial (8 layers upper)
Polynomial (8 layers lower)
0
500
1000
1500
2000
2500
0 100 200 300 400 500 600 700
Flow front position (mm)
T
i
m
e
(
s
e
c
)
11/14 layers upper
11/14 layers lower
Polynomial (14 layers upper)
Polynomial (14 layers lower)
Polynomial (11 layers upper)
Polynomial (11 layers lower)
T
i
m
e
(
s
)
T
i
m
e
(
s
)
- 163 -
Fig. 3. Time as a function of flow front position on both upper and lower faces of the preform.
Effect of thickness increase on infusion time.
Trendlines are polynomial curve fit.
Fig. 4. Time as a function of flow front position on both upper and lower faces of the preform.
Effect of thickness increase on infusion time.
Trendlines are polynomial curve fit.
0
200
400
600
800
1000
1200
1400
0 100 200 300 400 500 600 700
Flow front position (mm)
T
i
m
e
(
s
e
c
)
11/8 layers upper
11/8 layers lower
Polynomial (11 layers upper)
Polynomial (11 layers lower)
Polynomial (8 layers upper)
Polynomial (8 layers lower)
0
200
400
600
800
1000
1200
1400
1600
0 100 200 300 400 500 600 700
Flow front position (mm)
T
i
m
e
(
s
e
c
)
14/11 layers upper
14/11 layers lower
Polynomial (14 layers upper)
Polynomial (14 layers lower)
Polynomial (11 layers upper)
Polynomial (11 layers lower)
T
i
m
e
(
s
)
T
i
m
e
(
s
)
- 164 -
Appendix 7
Cost analysis of re-usable rubber membrane
The use of a new vacuum membrane in VI to replace the current distribution
medium and vacuum bag involves some cost changes and both current cost and
new bag cost have to be compared. They both involve a purchase and a labour cost.
The purchase cost involves the vacuum bag, distribution medium, springs, valves,
pump, reinforcement and resin. The new set-up requires an additional cost for the
manufacture of the vacuum membrane that needs to be cured to the shape of the
mould. The material has to be chosen, according to its elongation, chemical
resistance, cost and ageing. The aim is to produce a re-usable membrane made of
rubber or silicone so that the high manufacturing cost is compensated by the ability
to mould several components with the same bag. Cost savings made on eliminating
the actual distribution medium, vacuum bag and reducing resin consumption must
be compared with the cost involved to manufacture the new membrane, its life span
and the production rate required. The labour cost involves the lay-up of the preform
and consumables, closing of the mould, preparation of the resin mix, control of the
opening and closing of valves as well as part release and mould cleaning.
The cost analysis was related to an industrial application which is the manufacture
of the semi-trailer developed and produced during Roadlite. The chassis measured
13 m long by 2.5 m wide which gave a total surface of 31 m
2
(see Picture 1). Table
1 shows the details of the labour and purchase cost involved with the manufacture
of a prototype chassis and using the current set-up (distribution medium, bleed out
fabric and smooth vacuum bag). The cost takes into account all the steps of the
process from the purchase of all materials to the preparation of the mould, set-up of
the infusion and release of the part. This cost is compared with that of a chassis
manufactured with the new vacuum membrane. For this cost comparison, the cost
of the bag was based on the cost of silicone 1453DC6054 from Diatex. The
additionnal cost to cure the silicone to the shape of the mould was not taken into
account. Results show that the current set-up presents a higher labour cost than
purchase cost. In the case of the new membrane purchase cost is much higher than
labour cost due to the very high cost of the silicone bag itself. However since the
bag is a one-off pay material but used several times, it is more appropriate to look
at the cost as a function of the number of parts produced. Figure 1 shows a
- 165 -
comparison of the cost as a function of production rate to manufacture a chassis for
semi-trailer using either the current set-up or the new bag. Results show that the
new bag proves to be cost effective only if re-usable and becomes cheaper than the
conventionnal VI process if at least 12 parts are produced with the same bag.
Pic. 1. Roadlite composite chassis.
Table 1. Cost breakdown of the Roadlite trailer using current set-up or new bag
Purchase cost () Labour cost (30/hour)
Description Quantity
Current
set-up
New bag Current set-
up (hour)
New bag
(hour)
Trailer surface aspect
Gel coat (kg) 34 76.84 76.84 6 6
Tissue (kg) 1 11.67 11.67
Resin for tissue (kg) 21 33.60 33.60
14 14
Preform
Triaxial Formax FGE 117 (kg) 404 606.00 606.00
UD Formax FGE 122 (kg) 91 136.50 136.50
64 64
Bagging materials
Bleed out fabric (m
2
) 40 52.40 52.40 6 0
Distribution medium (m
2
) 40 44.80 0.00 8 0
Vacuum bag (m) 30 40.50 0.00 8 0
New bag (m
2
) 31 0.00 4495.00 8
Consumables
Wire coils (m) 80 50.40 0.00 0
Tubing (m) 60 37.80 0.00
6
0
Tacky tape (roll) 12 30.00 0.00 0
Resin system
Resin crystic 701PAX (kg) 150 240.00 240.00
Catalyst M50 (kg) 5.5 15.68 15.68
8 8
De-moulding 16 16
TOTAL () 1376.19 5667.69 4080 3480
- 166 -
0
50000
100000
150000
200000
250000
300000
0 10 20 30 40 50 60
Production rate (Parts)
C
o
s
t
(

)
Current set-up
New bag
Fig. 1. Cost of the current set-up and the new bag as a function of production rate.
- 167 -
Appendix 8
References in alphabetical order by author
1. S.G Advani et al. Flow and rheology in polymer composites manufacturing.
Newark, Elsevier, ISBN 0-444-89347-4 (Vol. 10), 1994. Chapter 12, p 465-
515.
2. H. M. Andersson, T.S. Lundstrom, B.R. Gebart, R. Langstrom. Flow enhancing
layers in the vacuum infusion process. Polymer Composites, Oct 2002, 23,
N
o
5, p 895-901.
3. M. Andersson, S. Lundstrom, B.R. Gebart, R. Langstrom. Development of
guidelines for the vacuum infusion process. 8
th
Fibre Reinforced Composite
Conference (FRC), Sep 13-15, 2000, Newcastle, UK, p. 113-120.
4. S. Bickerton, E.M Sozer, R.J Graham, S.G Advani. Fabric structure and mold
curvature effects on preform permeability and mold filling in the Resin Transfer
Molding process. Part 1: experiments. Composites Part A, 31, 2000, p 423-
438.
5. W.D. Brouwer, E. Van Herpt, M. Labordus. Vacuum injection moulding for large
structural applications. Composites Part A, 34, 2003, p. 551-558.
6. M.V Bruschke, S.G Advani. Flow of generalized Newtonian fluids across an
array of cylinders. Journal of Rheology, 37, 3, May/June 1993, p. 479-497.
7. J. P. Chick, C. D. Rudd, P. A. van Leeuwen, T. I. Frenay. Material
Characterization for Flow Modeling in Structural Reaction Injection Molding.
Polymer Composites, February 1996, 17, N
o
1, p. 124-136.
8. N. C. Correia. Analysis of the vacuum infusion moulding process. PhD thesis,
2004, University of Nottingham.
- 168 -
9. N.C Correia, F. Robitaille, A.C Long, C. D Rudd, P. Simacek, S.G Advani. Use of
resin transfer moulding simulation to predict flow, saturation and compaction
in the VARTM process. Journal of Fluids Engineering. ASME, 126 (2), 2004, p.
210-215.
10. A. Duarte. In-plane permeability of interleaved preforms for resin transfer
moulding.
11. M. Feiler, L. Ischtschuk. Vacuum assisted resin infusion (VARI): on the way to
serial production. 24
th
International SAMPE Europe Conference, April 2003,
Paris, France, p. 683-691.
12. B.R Gebart. Permeability of unidirectional reinforcements for RTM. Journal of
Composite Materials, 26, n8. 1992, p. 1100-1133.
13. A. Hammami, B.R. Gebart. Experimental investigation of the vacuum infusion
molding process. 8
th
European Conference on Composite Materials (ECCM-8),
June 1998, Naples, Italy, p. 577-584.
14. A. Hammami, B.R. Gebart. A model for the vacuum infusion molding process.
7
th
Fibre Reinforced Composite Conference (FRC), 1998, Newcastle, UK.
15. A. Hammami, B.R. Gebart. Model for vacuum infusion molding process.
Plastics, Rubber and Composites Processing and Application, 1998, 17, N4.
16. A. Hammami, B.R. Gebart. A model for the vacuum infusion moulding process.
7
th
Fibre Reinforced Composite Conference (FRC), 1998, Newcastle, UK, p.
136-145.
17. A. Hammami, R. Gauvin, F. Trochu. Modelling the edge effect in liquid
composites moulding. Composites Part A, 29, 1998, p. 603-609.
18. K. Han, S. Jians, C. Zhang, B. Wang. Flow modeling and simulation of SCRIMP
for composites manufacturing. Composites Part A, 31, 2000, p 79-86.
- 169 -
19. D. Heider, A. Karamitsos, J. W Gillespie, S. Walsh, E. Rigas, W. Spurgeon, W.
Roy. Experimental validation and optimisation of the FASTRAC process. 22
th
International SAMPE Europe Conference, April 2001, Paris, France, p 293-303.
20. A. Hoebergen, J. Holmberg. Vacuum infusion, ASM Handbook, 21, Composites,
Materials Park (OH), 2001, p. 501-515.
21. K-T. Hsiao, R. Mathur, S. G. Advani, J. W. Jr Gillespie, B. K. Fink. A closed
form solution for flow during the vacuum assisted resin transfer molding
process. Journal of manufacturing science and engineering, Aug 2000, 122, p
463-475.
22. P. Hubert, R. Byron Pipes, B.W. Grimsley. Variability analysis in vacuum
assisted resin transfer moulding. 23
th
International SAMPE Europe Conference,
April 2002, Paris, France, p. 415-426.
23. M.K. Kang, W.I. Lee, H.T. Hahn. Analysis of vacuum bag resin transfer
moulding process. Composites Part A, 32, 2001, p 1553-1560.
24. M. Koefoed, E. Lund, L. Lilleheden. Modeling and simulation of the VARTM
process for wind turbine blades. 10
th
European Conference on Composite
Materials (ECCM-10), June 3-7, 2002, Brugge, Belgium.
25. M. Labordus. Infusion of thick walled carbon epoxy structure. 22
th
International
SAMPE Europe Conference, April 2001, Paris, France.
26. W.S Leenders, P. Marissen, L. Drift, B. Boon. Special problems due to vacuum
injection of large ship structures. Flow Processes in Composite Materials
conference, 1999, Plymouth, England, p. 329-336.
27. J. Ni, S. Li, X. Sun, L.J. Lee. Mold filling analysis in vacuum assisted resin
transfer molding. Part II: SCRIMP based on grooves. Polymer Composites,
Dec. 1998, 19, N6, p. 818-829.
- 170 -
28. T. D. Papathanasiou, P. D. Lee. Morphological effects on the transverse
permeability of arrays of aligned fibers. Polymer Composites, April 1997, 18,
N
o
2, p. 242-253.
29. R.S. Parnas, A.J. Salem, T.A.K. Sadiq, H-P Wang, S.G Advani. The interaction
between micro- and macroscopic flow in RTM preforms. Composite structures,
27, 1994, p. 93-107.
30. R.S. Parnas, F.R. Phelan. The effect of heterogeneous porous media on mold
filling in resin trasnfer molding. SAMPE Quarterly, Jan 1991, 22, p. 53-60.
31. B. Qi, J. Raju, T. Kruckenberg, R. Stanning. A resin film infusion process for
manufacture of advanced composite structures. Composite Structures, 47,
1999, p 471-476.
32. A. Ragondet, N.C. Correia, F. Robitaille, A.C. Long, C.D. Rudd. Experimental
investigation on vacuum infusion process. 9
th
Fibre Reinforced Composites
conference (FRC), 2002, Newcastle, England, p. 437-445.
33. A. Ragondet, N.C. Correia, F. Robitaille, A.C. Long, C.D. Rudd. Experimental
investigation and modelling of the vacuum infusion process. Tenth European
Conference on Composite Materials (ECCM-10), June 3-7, 2002, Brugge,
Belgium.
34. F. Robitaille, R. Gauvin. Compaction of textile reinforcement for composites
manufacturing. I: review of experimental results. Polymer Composites 19, N
o
2, April 1998.
35. T.A.K. Sadiq, S.G. Advani, R.S. Parnas. Experimental investigation of
transverse flow through aligned cylinders. International Journal Multiphase
Flow, 21, N
o
5, 1995, p. 755-774.
36. T Searle, J Spooner, S Grove, R Cullen, S Davy. Manufacture of large
composite structures by vacuum resin infusion. 24
th
International SAMPE
Europe conference, April 2003, Paris, France, p 113-120.
- 171 -
37. W. H Seemann. Plastic transfer moulding techniques for the production of fibre
reinforced plastic structures. US patent No 4902215, filed March 1989.
38. P. Simacek, J. Lawrence, S. Advani. Numerical mold filling simulations of liquid
composite molding processes. Applications and current issues. 23
th
International SAMPE Europe Conference, April 2002, Paris, France, p. 137
148.
39. X. Song, A.C. Loos, B. Grimsley, R. Cano, P. Hubert. Modelling manufacture of
textile composites by VARTM. 6
th
International Conference on Textile
Composites (TexComp 6), Sept 2002, Philadelphia, USA.
40. X. Sun, S. Li, L.J. Lee. Mold filling analysis in vacuum assisted resin transfer
molding. Part I: SCRIMP based on a high permeable medium. SPE Technical
journals, Polymer Composites, Dec. 1998, 19, N6, p. 807-817.
41. M. J Tari, J-P Imbert, M. Y Lin, A. S Lavine, H. T Hahn. Analysis of resin
transfer moulding with high permeability layers. Journal of Manufacturing
Science and Engineering. Aug 1998, 120, p 609-616.
42. M. Thibaudeau, R.A. Shenoi. Understanding flow behaviour in the resin infusion
process. 10
th
European Conference on Composite Materials (ECCM-10), June 3-
7, 2002, Brugge, Belgium.
43. I. J. Verhaeghe. Composittrailer introducing composites in a conservative
environment. 10
th
European Conference on Composite Materials (ECCM-10),
June 3-7, 2002, Brugge, Belgium.
44. I. C. Visconti, A. Langella, M. Durante. Analysis of transversal permeability for
different types of glass fiber reinforcement. Applied Composite Materials,
March 2003, 10, N
o
2, p. 119-127.
45. G. Weijde, A. Brodsjo, J. Tyrrell, L. Van Rijn, M. Brooker. Development of low
cost composite fairings for the Ariane 5 engine thrust frame applying vacuum
- 172 -
assisted RTM. 23
th
International SAMPE Europe Conference, April 2002, Paris,
France, p 493-502.
46. C. Williams, J. Summerscales, S. Grove.Resin infusion under flexible tooling
(RIFT): a review. Composites Part A, 27, 1996, p 517-524.
47. C-H. Wu, T. J. Wang, L.J. Lee. Trans-plane fluid permeability measruement
and its applications in liquid composite molding. Polymer Composites, August
1994, 15, N
o
4, p. 289-298.

Вам также может понравиться