You are on page 1of 21

Advanced Composite Materials

ISSN: 0924-3046 (Print) 1568-5519 (Online) Journal homepage: https://www.tandfonline.com/loi/tacm20

Impregnation and resin flow analysis during


compression process for thermoplastic composite
production

Osuke Ishida, Junichi Kitada, Katsuhiko Nunotani & Kiyoshi Uzawa

To cite this article: Osuke Ishida, Junichi Kitada, Katsuhiko Nunotani & Kiyoshi Uzawa (2020):
Impregnation and resin flow analysis during compression process for thermoplastic composite
production, Advanced Composite Materials

To link to this article: https://doi.org/10.1080/09243046.2020.1752964

Published online: 19 Apr 2020.

Submit your article to this journal

Article views: 47

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tacm20
Advanced Composite Materials, 2020
https://doi.org/10.1080/09243046.2020.1752964

Impregnation and resin flow analysis during compression process for


thermoplastic composite production
Osuke Ishidaa*, Junichi Kitadab, Katsuhiko Nunotania and Kiyoshi Uzawaa
a
Innovative Composite Center, Kanazawa Institute of Technology, Hakusan, Japan; bEngineering
Group, Ipco K.k, Kobe, Japan
(Received 10 March 2020; accepted 3 April 2020)

Impregnation and in-plane resin flow of film-stacking material under compression


process were investigated. The understanding of this phenomenon is the key to
optimize the impregnation process under pressure rollers in a double-belt press for
organo-sheet production. The stack of carbon fiber woven fabrics and viscous fluid
were compressed in a flat open-edge mold under various controlled conditions using
a testing machine. The impregnation, in-plane flow, and pressure profile were char-
acterized from the experimental observations. Furthermore, a theoretical model was
developed to describe the impregnation and in-plane flow behavior. The model
prediction showed decent agreement with the experimental data. The developed
model will be useful to optimize the process parameters of double belt press process.
Keywords: thermoplastic composites; impregnation; resin flow; process modelling;
double belt press

1. Introduction
Carbon fiber reinforced thermoplastics are gaining remarkable interest because they have
advantages such as short cycle times, toughness, and recyclability. In recent years,
various fundamental studies have been conducted on the application of this material
[1–3]. Thermoplastic composite sheet, also known as organo-sheet, is a promising
material for industrial and automobile parts because it enables a short cycle production
by stamp forming and welding [4–8]. Normally, organo-sheet can be produced by
compression molding using fiber reinforcements and thermoplastic resin. The key aims
are to achieve good impregnation quality and low cost. However, it is difficult to achieve
these properties together because the impregnation is difficult due to the high melt
viscosity. Therefore, the impregnation technology has been a subject of research interest
for thermoplastic composites [9–12].
A double-belt press is one of the solutions for organo-sheet production. The machine
consists of heat and pressure modules for impregnation and a subsequent cooling module
for solidification (Figure 1), enabling a continuous compression molding process.
Several investigations have been reported about this system [13–18]. There are two
types of press systems for a double-belt press, the isobaric press and the roller press
(Figure 2). The isobaric press provides constant and uniform pressure for the impregna-
tion, and the process parameters can be easily controlled by a machine setup. However,
the roller press is more suitable for the impregnation process in a double-belt press

*Corresponding author. Email: o-ishida@neptune.kanazawa-it.ac.jp

© 2020 Japan Society for Composite Materials, Korean Society for Composite Materials and Informa UK Limited, trading as
Taylor & Francis Group
2 O. Ishida et al.

Figure 1. Schematic of the fixed rollers double-belt press for organo-sheet production.

Figure 2. Difference of the isobaric press system and the roller press system.

because the machine cost is relatively lower than isobaric press. Additionally, different
types of materials can be applied, such as discontinuous reinforcements and thick
composites. The material is compressed along the steel belt deflection under a roller
(Figure 3). As there is a pressure gradient between the outside of a roller and the region
underneath it, not only fabric impregnation but in-plane resin flow occurs as illustrated in
Figure 4. This phenomenon is prominent while using a film-stacking material. The
applied pressure is not constant under a roller, which is determined by the balance of
the impregnation and in-plane flow under compression. The difficulty of optimizing the
process parameters is a critical issue for the composite manufacturing process using
a double belt press. Thus, the understanding of the resin flow mechanism under a roller is
very important.
Resin impregnation process during compression molding has been studied by many
researchers [19–24]. In these studies, the resin flow through fiber network has been
Advanced Composite Materials 3

Figure 3. Material compression behaviour under a roller in the double belt press (in the case of
film-stacking material).

Figure 4. Impregnation and in-plane flow behavior under a roller in the double belt press.

modelled based on Darcy’s law. For example, Gutowski et al. [19] evaluated the
transverse permeability of fiber bundles based on fiber elastic deformation model.
Michaud et al. [20] developed the impregnation model of glass mat thermoplastic
(GMT) based on local resin flow in compressible porous media. Kobayashi et al. [21]
studied the impregnation behavior of Micro-Braided-Yarn based on the ellipsoidal model
of fiber yarn and evaluated the effects of the yarn conditions. For the compression resin
transfer molding (CRTM) process, Merrote et al. [22] modelled the resin in-plane flow in
fiber preform during compression process. Their works have successfully explained the
resin flow behavior through fiber reinforcements. However, these analyses cannot
describe the combined process of the fiber bundle impregnation and resin in-plane
flow of film stacking material. The aim of our study is to clarify this combined
impregnation and in-plane flow behavior.
In this study, model experiments were conducted to study this flow behavior. The
stack of carbon fiber woven fabrics and viscous fluid were compressed using a testing
machine equipped with a flat mold under several process conditions. The fabric
4 O. Ishida et al.

impregnation, in-plane flow, and pressure profiles were characterized from the experi-
mental observations. Furthermore, a theoretical model was developed to describe the
impregnation and in-plane flow behavior. The predicted result was compared with the
experimental data.

2. Materials and methods


2.1. Materials
Carbon fiber plain woven fabric (T700SC-12 K, 3.27 yarns/25 mm, 210 g/m2, Toray co.
ltd.) was used as the reinforcement. Polyamide 6 film (CM1006, thickness 100 μm,
Toray co., ltd.) was used as the matrix resin for the compression molding experiments.
Viscous fluid (JS160000, Nippon Grease Co., Ltd.) was used as the matrix resin for the
impregnation and in-plane flow experiments using a testing machine. The fluid viscosity
was 151.5 Pa·s at 20 ºC, which was similar to the melt viscosity of polyamide 6. The
mold thermal expansion could negatively influence the accuracy of measuring the
material thickness reduction, and therefore, the viscous fluid was used in this experiment.
By using the fluid, the impregnation and in-plane flow behavior could be studied without
heating the mold.

2.2. Compression molding experiment


Five layers of carbon fiber woven fabrics and six layers of polyamide 6 films were
stacked alternately. The material size was 150 mm × 150 mm. The compression molding
experiments were conducted using a hot press machine (Shinto Metal Industries, Ltd.).
The material was placed on the heated plate and heated up to a temperature of 235 ºC,
and then, compressed at a pressure of 0.78 MPa for a holding time varying from 6 s to
120 s. Afterwards, the material was cooled down to a temperature less than 60 ºC. The
specimens were machined from the composites. The cross-section was polished and
observed by microscopy.

2.3. Impregnation and in-plane flow experiment using a testing machine


2.3.1. Experimental equipment
A metal mold (cavity size: 40 mm × 40 mm) was mounted on a fixture that was
connected to a load cell (5 kN) and a testing machine (AG-Xplus, Shimazu Co., Ltd.).
The mold setup was available with shear-edge and open-edge configurations by attaching
the side walls. The experimental system is illustrated in Figure 5. The cross-head position
and the force were recorded at a time step of 0.1 s. The fixture deformed slightly due to
the applied force during the experiment. Therefore, the deformation values were mea-
sured as a function of the force by conducting a blank experiment and then subtracted
from the measured values in the experiments.

2.3.2. Fabric compression measurement


Five layers of carbon fiber woven fabrics were compressed in the mold at a cross-head
speed of 0.5 mm/min and 1 mm/min. The resulting stress and fiber volume fraction were
calculated from the position and force of the testing machine. The average value was
obtained from five experiments.
Advanced Composite Materials 5

Figure 5. Experimental equipment using a testing machine. The mold setup is available with (a)
shear-edge and (b) open-edge configurations by attaching the side walls.

2.3.3 Impregnation and in-plane flow experiments


In the first step, fabric impregnation experiments were conducted. The viscous fluid was
applied manually on both surfaces of the carbon fiber woven fabric with an equivalent
thickness of 50 μm using a scraper. The applied amount was controlled by the weight.
Five layers of materials were compressed in a shear-edge mold at the cross-head speeds
of 0.5 mm/min and 1 mm/min until the applied force reached 2200 N (corresponding to
1.38 MPa). In addition, the same experiment was performed at the controlled forces of
640 N (0.4 MPa) and 1280 N (0.8 MPa) for a holding time of 60 s. The material
thickness reduction and force change were monitored during the process.
In the second step, in-plane flow experiments were conducted. The viscous fluid with
an equivalent thickness of 200 μm was placed in the open-edge mold and compressed at
the cross-head speeds of 0.5 mm/min and 1 mm/min until the applied force reached
2200 N (1.38 MPa). The material thickness reduction and force change were monitored
during the process.
In the third step, the combined experiments of fabric impregnation and in-plane flow
were conducted. The same materials as used in the fabric impregnation experiments were
placed in the open-edge mold and compressed at a controlled force of 640 N (0.4 MPa)
for a holding time of 60 s. Additionally, the combined experiments of fabric impregna-
tion and in-plane flow were performed at the cross-head speeds of 0.5 mm/min and
1 mm/min until the applied force reached 2200 N (1.38 MPa). After the experiment was
finished, the outflow of fluid at the mold edge was absorbed using Kimwipes and the
amount was weighed.

2.3.4 In-plane permeability measurement


The in-plane permeability of carbon fiber woven fabrics was measured using a dedicated
machine (EASYPERM®, IS Group). Five layers of the fabrics were compressed in the
thickness direction to obtain the desired fiber volume fraction vf = 0.45, 0.5, and 0.55 in the
chamber. Canola oil (The Nisshin OilliO Group, Ltd.) was injected into a small hole
punched at the center of the stack. The perimeter was maintained under atmospheric
pressure. The fabrics were stacked at identical orientation (x-axis: warp direction, y-axis:
weft direction). After the radial flow became stable, the pressure-drop along a certain length
6 O. Ishida et al.

in the two directions (x, y) were measured with an embedded pressure sensor. The
permeabilities were calculated based on Darcy’s equation in polar coordinates from the
measured pressure drops. The fluid viscosity was measured to be 44 mPa·s and the injection
pressure was 0.28 MPa. The average value was obtained from four measurements.

3. Results and discussion


3.1. Microscopy of impregnation process
Figure 6 shows the cross-sectional micrographs of composites made by the compression
molding process at different holding times. It is found that the resin instantly fills up the
space outside the fiber bundles at a holding time of 6 s. Subsequently, the resin starts to
infiltrate into the fiber bundles. The fiber arrangement at the interface between the resin
layer and fiber bundle is non-uniform, which might cause the waviness of the flow front.
The flow front advances inside the fiber bundle with holding time. The flow seems to
almost be in one direction due to the high aspect ratio of the bundle cross-section. The
fabric relaxation can be hardly observed during the impregnation process. The possible
reason is that the relaxation is very slow due to the high resin viscosity and the effect is
limited in the range of the experimental conditions.

3.2. Fabric compression behavior


The compression behavior of the stack of dry carbon fiber woven fabrics is shown in Figure
7. The compression model has been investigated in several studies and it is known that the
fabric compression behavior is viscoelastic [24,25]. The slight difference of the stress curve
at the cross-head speeds of 1 mm/min and 0.5 mm/min might be due to the viscoelastic
effect. However, this is not clear when compared with the variation of the values. The
relationship between the applied stress P and the fiber volume fraction vf has been approxi-
mated with a power law as shown in Equation (1). It does not fit either curve perfectly, but
fits with reasonable accuracy. This approximation is used for the impregnation analysis in the
following discussion.

vf ¼ 0:560  P0:045 (1)

3.3. Impregnation analysis using a testing machine


Figure 8 shows the measured force change and material thickness reduction with time
using a testing machine equipped with a shear-edge mold. The time starts when the upper
mold contacts the material top surface. The thickness reduction was calculated by sub-
tracting the fixture deformations from the measured cross-head displacements over the
range of forces, therefore, the rates of thickness reduction are not constant in Figure 8(d).
The material thickness reduction can be related to the fabric impregnation. The mass
conservation in the mold during the impregnation process is described according to the
following equation.

Vf  V i f
þ V i f þ Vr ¼ Vm (2)
vf
Advanced Composite Materials 7

Figure 6. Cross-sectional micrographs of composites made by the compression molding process


at different holding times: (a) 6 s-overview, (b) 6 s, (c) 60 s, (d) 120 s.
8 O. Ishida et al.

0.6

Fiber volume fraction


0.55

0.5

0.45 0.5 mm/min


1 mm/min
Fitting
0.4
0 0.5 1 1.5
Stress (MPa)
Figure 7. Compression behavior of carbon fiber woven fabrics at different compression speeds.

where Vf is the fabric solid volume, Vif is the impregnated part of the fabric solid volume, Vr
is the fluid volume, and Vm is the mold cavity volume. In this equation, (Vf – Vif) indicates
the solid volume of fibers in the unimpregnated part. The fibers in the unimpregnated part
are compressed by the stress equivalent to resin pressure during the impregnation process.
Thus, the fiber volume fraction vf in the unimpregnated part can be given by Equation (1).
By dividing (Vf – V if) by vf, one can obtain the bulk volume of unimpregnated part that
consists of the unimpregnated fibers and voids. Finally, the left-hand side of Equation (2)
comprises of the volume of unimpregnated fibers and voids (the first term), the volume of
impregnated fibers (the second term) and the volume of resin (the third term).
Based on the cross-sectional observation in Figure 6, the impregnation process can be
modeled as illustrated in Figure 9. The resin fills up the space outside the fiber bundles at
the initial stage, and then it starts to infiltrate into the fiber bundles in one direction. The
time increment of the impregnation length l in the fiber bundle can be expressed
according to the following equation.

ΔV i f
Δl ¼ (3)
10Avf

where A is the base area of the mold cavity. There are total ten impregnation flows with both
surfaces of the five layers of fabrics, thus, which explains the division by 10 in Equation (3).
The impregnation behavior can be described by Darcy’s law [19–24,26-30]. In this study, the
equation is defined as follows.

  Δl Kb P
1  vf ¼ (4)
Δt μ l

where P is the resin pressure, t is the time, and μ is the resin viscosity. We refer to Kb as
the apparent fiber bundle permeability in this study. The degree of impregnation is
defined as follows.
Advanced Composite Materials 9

(a) 1400
1200
1000

Force (N)
800
600
400
640 N
200
1280 N
0
0 20 40 60
Time (s)

(b) 2500
0.5 mm/min
2000 1 mm/min
Force (N)

1500

1000

500

0
0 40 80 120 160
Time (s)

(c) 2
640 N
1.8 1280 N
Thickness (mm)

1.6

1.4

1.2

1
0 20 40 60
Time (s)

(d) 2
0.5 mm/min
1.8 1 mm/min
Thickness (mm)

1.6

1.4

1.2

1
0 40 80 120 160
Time (s)

Figure 8. Experimental results of (a), (b) applied force change and (c), (d) material thickness
reduction under different process conditions using the testing machine.
10 O. Ishida et al.

Figure 9. Impregnation model of fiber bundles of woven fabric.

V if
α¼ (5)
Vf

Using Equations (1), (2) and (5), one can calculate the degree of impregnation. Using
Equations (1)–(4), one can calculate the apparent bundle permeability. These are obtained at
a time step of Δt from the experimental data. The calculated results are shown in Figure 10.
The relationship between the impregnation rate, impregnation length, and applied pressure
can be explained based on Equation (4). If the applied pressure is constant, the impregnation
rate will decrease with increasing impregnation. On the other hand, if the impregnation rate
is constant, the applied pressure will increase with increasing impregnation. This explana-
tion is consistent with the experimental results of the force profiles in Figure 8(a,b) and the
degree of impregnation in Figure 10(c,d). It can be observed that the apparent fiber bundle
permeability decreases with time. This may be attributed to the following effects. The filling
process outside the fiber bundles can cause a very high permeability in the initial stage.
Fiber alignment at the interface between the resin layer and fiber bundle is non-uniform as
shown in Figure 6. This causes a higher permeability compared to the core region of the
fiber bundle. Another effect is air entrapment inside the fiber bundles, which especially
occurs in woven fabric architectures [14,26]. The entrapped air generates internal pressure
due to the advancing flow front. Consequently, the apparent fiber bundle permeability
decreases with time. Based on these considerations, we thought that the apparent fiber
bundle permeability could be described as a function of the degree of impregnation. The
result is shown in Figure 11. The resulting curves are similar across the different process
conditions. Slight differences between the curves may be attributed to the difference of the
fiber volume fractions, which vary based on the different pressure conditions as suggested
by the fabric compression behavior in Figure 7. It is known that the transverse permeability
of fiber bundles depends on the fiber volume fraction [19–21]. However, the permeability
difference is not significant in this experimental range when it is compared with the
variation of the fiber volume fraction. The curves are approximated using a Gaussian
function as shown in Equation (6). This approximation is used for the impregnation
simulation in the following discussion.

 
14 α2
Kb ¼ 4:3  10  exp  (6)
0:28
Advanced Composite Materials 11

(a) 4E-14
640 N
1280 N

Bundle permeability (m2)


3E-14

2E-14

1E-14

0
0 20 40 60
Time (s)

(b) 4E-14
0.5 mm/min
1 mm/min
Bundle permeability (m2)

3E-14

2E-14

1E-14

0
0 40 80 120 160
Time (s)

(c) 1

0.8
Degree of impregnation

0.6

0.4

0.2 640 N
1280 N
0
0 20 40 60
Time (s)

(d) 1

0.8
Degree of impregnation

0.6

0.4

0.2 0.5 mm/min


1 mm/min
0
0 40 80 120 160
Time (s)

Figure 10. (a), (b) Apparent fiber bundle permeability and (c), (d) the degree of impregnation
calculated from the experimental data under different process conditions.
12 O. Ishida et al.

4E-14
640 N
1280 N

Bundle permeability (m2)


3E-14 0.5 mm/min
1 mm/min
Fitting
2E-14

1E-14

0
0.1 0.3 0.5 0.7 0.9
Degree of impregnation
Figure 11. Relationship between the apparent fiber bundle permeability and the degree of
impregnation under different process conditions.

3.4. In-plane flow analysis using a testing machine


The in-plane flow experiments have been conducted using a testing machine equipped
with an open-edge mold. In these experiments, just the one-dimensional flows in
opposite directions were created in the open-edge mold as shown in Figure 5. The
flow can be presumed as Stokes flow because the Reynolds number is very low due to
the low velocity and high viscosity. The flow behavior of the thin viscous layer between
the two mold surfaces can be expressed by the Reynolds equation [31,32], which is
derived from the Navier-Stokes equation with several assumptions such as thin film
geometry, negligible inertial and body force. The one-dimensional flow can be expressed
under the above assumptions as the following equation.

@P @2u
¼μ 2 (7)
@x @z

where u is the velocity in the x-direction (in-plane flow direction) and z is the thickness
direction. The no-slip boundary conditions are:

uð0Þ ¼ u1 and uðhÞ ¼ u2

where h is the fluid thickness, u1 and u2 are the velocities in the x-direction at the top
and bottom surfaces. By integrating Equation (7) twice with the above boundary condi-
tions, we get:

1 @P  2   z z
u¼ z  zh þ 1  u1 þ u2 (8)
2μ @x h h
Advanced Composite Materials 13

In this experiment, u1 and u2 are zero at the mold surfaces. There is no flow in the width
direction, thus, the continuity equation can be formulated as follows.

@u @v
þ ¼0 (9)
@x @z

where v is the velocity in the z-direction. By substituting Equation (8) to (9) and by
integrating across the thickness from z = 0 to z = h, one can obtain the following
equation.
 
@ h3 @P
V ¼ (10)
@x 12μ @x

where V is the compression speed. The pressure field inside the mold can be determined
by applying the following two boundary conditions at the center position (symmetry
boundary condition) and the edge position Le in the mold cavity.
 
@P
¼ 0 and Px¼Le ¼ 0
@x x¼0

Thus, the fluid pressure can be calculated as follows.

6μV  2 
Pð x Þ ¼ Le  x2 (11)
h3

The force acting on the mold plate F can be obtained by integrating the fluid pressure on
the surface.
ð Le
8wμLe 3 V
F ¼ 2w Pð xÞdx ¼ (12)
0 h3

where w is the width of the fluid. The applied force F can be calculated at each time step
using Equation (12) when the compression speed is given.
The comparison of the experimental and prediction force profiles is shown in Figure 12.
The theoretical model achieves a good fit with the data. A slight discrepancy might have been
caused due to the theoretical assumption of no-slip boundary condition at the mold surface.

3.5. Result of in-plane permeability measurement


Figure 13 shows the measured permeability of the carbon fiber woven fabrics in the in-
plane direction using the dedicated machine. The result indicates the permeability in the
x-direction that corresponds to the in-plane flow direction of the stacked fabrics in the
open-edge mold. The permeability Ki decreases as the fiber volume fraction increases.
There is a large variation of the values at low vf = 0.45. The reason may be that the fabric
could not be fixed completely in the chamber. The curve is approximated using a power
law as shown in Equation (13). This approximation is used for the in-plane flow
simulation in the fabric layer in the following discussion.
14 O. Ishida et al.

2500

2000

Load (N)
1500

1000

500 0.5 mm/min


1 mm/min
0
0 20 40 60 80
Time (s)
Figure 12. Applied force profiles in the in-plane flow experiments at different compression
speeds: solid lines are the prediction results, dotted lines are experimental data.

1.6E-10
Experiment
Fitting
In-plane permeability (m2)

1.2E-10

8E-11

4E-11

0
0.4 0.45 0.5 0.55 0.6
Fiber volume fraction
Figure 13. Measured in-plane permeability of the carbon fiber woven fabrics.

Ki ¼ 1014  vf 11:5 (13)

3.6. Analysis of the combination of impregnation and in-plane flow using a testing
machine
The combination experiments of impregnation and in-plane flow have been conducted
using a testing machine equipped with an open-edge mold. When the mold edge is under
atmospheric pressure, there is a pressure gradient between the center position and the
edge in the mold, which causes in-plane flow in the resin layer and fabric layer.
Advanced Composite Materials 15

Simultaneously, the fiber bundle impregnation is caused by a pressure gradient between


the resin layer and the un-impregnated area of the fiber bundle. In order to develop the
flow process model, we subdivided the process area in the in-plane direction and
considered the conservation of mass flow in each area. Figure 14 illustrates this model
that consists of a half layer of fabric and one resin layer. The resin pressure is described
as a variable for each area and the equations of the resin flows are formulated based on
the previous discussion. The fiber bundle impregnation flow rate Qb can be expressed as
the following equation based on the discussion in Section 3.3.

Qb   Δl Kb P
¼ 1  vf ¼ (14)
wd Δt μ l
where w is the width of the in-plane flow, d is the in-plane length of each area, l is the
impregnation length in the fiber bundle, and P is the resin pressure.
The in-plane flow rate in fabric layer Qf can be expressed as Darcy flow. The in-plane
permeability Ki was previously obtained in Section 3.5. The equation is defined as follows.

Qf Ki ΔP
¼ (15)
wl μ 2d

where ΔP indicates the pressure difference between the adjacent areas. There is
a pressure gradient from the resin pressure in the resin layer P to the atmospheric
pressure in the unimpregnated part of a fiber bundle due to the fiber bundle impregnation
as illustrated in Figure 9. We took the mean pressure (P/2) to define the resin pressure

Compression Center line


Resin
Mold Fabric

Qr P1 P2 P3 P4 P5 h In-plane flow
Outflow
Qf Fiber bundle
l impregnation
Qb
z
d Mold

x
Resin pressure

Figure 14. The combination model of impregnation and in-plane flow during compression
process.
16 O. Ishida et al.

inside the fabric in Equation (15) for simplicity, which explains the division by two in the
equation.
The in-plane flow rate in the resin layer Qr can be expressed as a thin-layer Stokes
flow between the fabric surfaces based on the discussion in Section 3.4. It can be
calculated by integrating Equation (8) across the resin layer thickness h. It has been
reported that slippage occurs at the interface between the fluid velocity in the resin layer
and Darcian velocity in the fabric layer, however, the effect can be negligible due to the
small transition distance [33]. Thus, the fluid velocity at the fabric surface is assumed to
be the same as that of the in-plane flow velocity in the fabric layer. Consequently, the
flow can be expressed as:

wh3 ΔP h
Qr ¼ þ Qf (16)
12μ d l

Finally, the conservation of mass flow in each area is expressed as:

ΔVr
¼ Qrin  Qrout þ Qfin  Qfout (17)
Δt

where ΔVr/Δt is the resin volume change with time in each area, which is determined by
the inflow rate Qrin and outflow rate Qrout in the resin layer, and the inflow rate Qfin and
outflow rate Qfout in the fabric layer. When Equation (2) and Equations (14)-(17) are
combined, one can formulate the system of equations for the mass conservation of the
whole area. Using the solver tool of Microsoft Excel, the system of equations has been
solved at a time step of 1 s. Consequently, we obtained the resin pressure in each area
and the compression speed at each time step. Figure 15 shows the compression speed
obtained from the model and experiment under a controlled force of 640 N (0.4 MPa).
Figure 16 shows the applied force profiles obtained from the model and experiment at the
controlled compression speeds of 0.5 mm/min and 1 mm/min. It is found that the

0.005
Model
Compression speed (mm/s)

0.004 Experiment

0.003

0.002

0.001

0
0 20 40 60
Time (s)
Figure 15. Comparison of compression speed obtained from the model and experiment under
a controlled force of 640 N (0.4 MPa).
Advanced Composite Materials 17

experimental data is well represented by the theoretical model. The degree of impreg-
nation obtained from the model and experiment is shown in Figure 17. The result shows
reasonable agreement. The values of in-plane outflow obtained from the model and
experiment are shown in Figure 18. The values of outflow are less than one-tenth of
the values of total fiber bundle impregnation (Σ Qb) and there are differences between the
predicted and experimental values. The possible causes are the variations of the experi-
mental setup and the inaccuracies of the approximation Equations (1), (6), and (13) as
discussed before.

2500
0.5 mm/min
1 mm/min
2000
Load (N)

1500

1000

500

0
0 20 40 60 80 100 120 140
Time (s)
Figure 16. Comparison of the force profiles obtained from the model and experiments under
controlled compression speeds of 0.5 mm/min and 1 mm/min: the solid line is the prediction, the
dotted line is experimental data.

1
Model Experiment
0.8
Degree of impregnation

0.6

0.4

0.2

0
(a) Speed: (b) Speed: (c) Force:
0.5 mm/min 1 mm/min 640 N
Figure 17. Comparison of the degree of impregnation obtained from the model and experiment
under different compression conditions: (a) controlled speed of 0.5 mm/min, (b) controlled speed
of 1 mm/min, (c) controlled force of 640 N.
18 O. Ishida et al.

50
Model Experiment
40

In-plane outflow (mm3)


30

20

10

0
(a) Speed: (b) Speed: (c) Force:
0.5 mm/min 1 mm/min 640 N
Figure 18. Comparison of the values of in-plane outflow obtained from the model and experi-
ment under different compression conditions: (a) controlled speed of 0.5 mm/min, (b) controlled
speed of 1 mm/min, (c) controlled force of 640 N.

Although a few improvements are possible, the theoretical model explains well, the
impregnation and in-plane flow of the film-stacking material under compression process.
In the future, this model will be applied to investigate the impregnation behavior under
rollers in a double-belt press process and optimize the process parameters.

4. Conclusions
In this study, the impregnation and in-plane flow of the film-stacking material under
compression process have been investigated based on experiments and a theoretical
model. The motivation of this study is to understand the impregnation behavior under
pressure rollers in a double-belt press for organo-sheet production. As there is a pressure
gradient between the outside of the rollers and the region underneath it, not only fabric
impregnation, but also in-plane resin flow occurs during the compression process. The
applied pressure is determined by these combination flows.
At first, the impregnation process has been investigated by the cross-section micro-
scopy of composites fabricated by compression molding. Subsequently, impregnation
experiments have been conducted using a testing machine equipped with a shear-edge
mold in several compression conditions. Consequently, the fiber bundle impregnation has
been characterized based on Darcy’s law. For the next step, the in-plane flow behavior
has been studied using a testing machine equipped with an open-edge mold. The
prediction model based on thin-layer Stokes flow has shown good agreement with the
experimental data. Finally, the combination behavior of fiber bundle impregnation and
in-plane flow has been studied through similar experiments. The experimental data has
been well represented by the prediction model, although a few improvements are
possible. In the future, the developed model will be applied to study the double-belt
press process in order to optimize the process parameters.
Advanced Composite Materials 19

Disclosure statement
No potential conflict of interest was reported by the authors.

Funding
This work was supported by the JST COI Grant Number [JPMJCE1315].

References
[1] Taketa I, Kalinka G, Gorbatikh L, et al. Influence of cooling rate on the properties of carbon
fiber unidirectional composites with polypropylene, polyamide 6, and polyphenylene sulfide
matrices. Adv Compos Mater. 2020;29:101–113.
[2] Wei H, Nagatsuka W, Ohsawa I, et al. Influence of small amount of glass fibers on mechanical
properties of discontinuous recycled carbon fiber-reinforced thermoplastics. Adv Compos
Mater. 2019;28:321–334.
[3] Guo Q, Ohsawa I, Takahashi J. L-shaped structures made from ultrathin chopped carbon fiber
tape reinforced thermoplastics: delamination behavior and optimization. Adv Compos Mater.
2019;28:479–489.
[4] Hou M. Stamp forming of continuous glass fiber reinforced polypropylene. Compos Part A.
1997;28A:695–702.
[5] Vieille B, Albouy W, Chevalier L, et al. About the influence of stamping on
thermoplastic-based composites for aeronautical applications. Compos Part B.
2013;45:821–834.
[6] Wang P, Hamila N, Boisse P. Thermoforming simulation of multilayer composites with
continuous fibres and thermoplastic matrix. Compos Part B. 2013;52:127–136.
[7] Yousefpour A, Hojjtai M, Immarigeon J-P. Fusion bonding/welding of thermoplastic
composites. Compos Part A. 2006;37(10):1638–1651.
[8] Shirai T, Uzawa K. Relationship between tensile properties and fiber orientation after press
forming of discontinuous carbon fiber reinforced thermoplastic composite. 16th Japan
International SAMPE Symposium; 2019 Sep 2–4;Tokyo, Japan. p. 3A04.
[9] Gibson AG, Månson J-A. Impregnation technology for thermoplastic matrix composites.
Compos Manuf. 1992;3(4):223–233.
[10] Khondker OA, Ishiaku US, Nakai A, et al. A novel processing technique for thermoplastic
manufacturing of unidirectional composites reinforced with jute yarns. Compos Part A.
2006;37:2274–2284.
[11] Nishida N, Nunotani K, Uzawa K. In situ-polymerizing thermoplastic epoxy resin which
enables molding processes corresponding to various forms of thermoplastic composites. 18th
European conference on composite materials; 2018 June 24-28; Athens, Greece. P-5.19.
[12] Ishida O, Uzawa K, Kinpara I. Production method of carbon fiber woven fabric/PA6 compo-
site using PA6 solution impregnation process. Reinf Plast. 2016;62(8): 318–323. Japanese.
[13] Wang X, Mayer C, Neitzel M. Some issues on impregnation in manufacturing of thermo-
plastic composites by using a double belt press. Polym Compos. 1997;18(6):701–710.
[14] Mayer C, Wang X, Neitzel M. Macro-and micro-impregnation phenomena in continuous
manufacturing of fabric reinforced thermoplastic composites. Compos Part A.
1998;29:783–793.
[15] Liu D, Zhu Y, Ding J, et al. Experimental investigation of carbon fiber reinforced poly
(phenylene sulfide) composites prepared using a double-belt press. Compos Part B.
2015;77:363–370.
[16] Trende A, Astrom BT, Woginger A, et al. Modeling of heat transfer in thermoplastic
composites manufacturing: double-belt press lamination. Compos Part A.
1999;30:935–943.
[17] Ishida O, Fukushima G, Kitada J, et al. Continuous impregnation process by fixed rollers
double belt press. 17th European conference on composite materials; 2016 June 26-30;
Munich, Germany. STG-5.09-04.
[18] Ishida O, Kitada J, Uzawa K. Investigation of continuous impregnation process by fixed
rollers double belt press. Mater Syst. 2017;35: 53–59. Japanese.
20 O. Ishida et al.

[19] Gutowski TG, Cai Z, Bauer S, et al. Consolidation experiments for laminate composites.
J Compos Mater. 1987;21:650–669.
[20] Michaud V, Manson J-AE. Impregnation of compressible fiber mats with a thermoplastic
resin. Part I: theory. J Compos Mater. 2001;35(13):1150–1173.
[21] Kobayashi S, Tsukada T, Morimoto T. Resin impregnation behavior in carbon fiber reinforced
polyamide 6 composite: effects of yarn thickness, fabric lamination and sizing agent. Compos
Part A. 2017;101:283–289.
[22] Merotte J, Simacek P, Advani SG. Flow analysis during compression of partially impregnated
fiber preform under controlled force. Compos Sci Technol. 2010;70:725–733.
[23] Ye L, Klinkmüller V, Friedrich K. Impregnation and consolidation in composites made of GF/
PP powder impregnated bundles. J Thermoplast Compos Mater. 1992;5:32–48.
[24] Lekakou C, Johafu MAKB, Bader MG. Compressibility and flow permeability of
two-dimensional woven reinforcements in the processing of composites. Polym Compos.
1996;17(5):666–672.
[25] Yenilmez B, Caglar B, Sozer EM. Pressure-controlled compaction characterization of fiber
preforms suitable for viscoelastic modelling in the vacuum infusion process. J Compos Mater.
2017;51(9):1209–1224.
[26] Foley ME, Gillespie JW Jr. Modeling the effect of fiber diameter and fiber bundle count on
tow impregnation during liquid molding processes. J Compos Mater. 2005;39(12):1045–1065.
[27] Murphy CS, Simacek P, Advani SG, et al. A model for thermoplastic melt impregnation of
fiber bundles during consolidation of powder-impregnated continuous fiber composites.
Compos Part A. 2010;41:93–100.
[28] Koubaa S, Le Corre S, Burtin C. Thermoplastic pultrusion process: modeling and optimal
conditions for fibers impregnation. J Reinf Plast Comp. 2013;32(17):1285–1294.
[29] Jespersen ST, Wakeman MD, Michaud V, et al. Film stacking impregnation model for a novel
net shape thermoplastic composite preforming process. Compos Sci Technol.
2008;68:1822–1830.
[30] Studer J, Dransfeld C, Cano JJ, et al. Effect of fabric architecture, compaction and perme-
ability on through thickness thermoplastic melt impregnation. Compos Part A.
2019;122:45–53.
[31] de Vicente J, Ruiz-López JA, Andablo-Reyes E, et al. Squeeze flow magnetorheology.
J Rheol. 2011;55(4):753–779.
[32] Ren F, Yu Y, Yang J, et al. A mathematical model for continuous fiber reinforced thermo-
plastic composite in melt impregnation. Appl Compos Mater. 2017;24:675–690.
[33] Hammami A, Gauvin R, Trochu F. Modeling the edge effect of liquid composites molding.
Compos Part A. 1998;29A:603–609.

You might also like