Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Composites: Part A 141 (2021) 106203

Contents lists available at ScienceDirect

Composites Part A
journal homepage: www.elsevier.com/locate/compositesa

Concurrent characterization of through-thickness permeability and


compaction of fiber reinforcements
M.A Kabachi *, L. Stettler , S. Arreguin , P. Ermanni
Laboratory of Composite Materials and Adaptive Structures, ETH Zurich, 8092 Zurich, Switzerland

A R T I C L E I N F O A B S T R A C T

Keywords: Flow-induced fiber-bed compaction has considerable effects on through-thickness permeability due to resulting
Through-thickness permeability changes in fiber-volume-fractions. This study introduces a novel method for the simultaneous characterization of
Fiber-bed compaction compaction and unsaturated out-of-plane permeability of engineering textiles in one single experiment. The
Unsaturated permeability
measurement technique relies on visually tracking flow-front positions and compaction during through-thickness
impregnation using a camera and a rigid tool equipped with a pressure sensor and a transparent window. This
strategy allows flow-induced fiber-bed compaction to be considered in unsaturated permeability measurements,
providing permeability at different fiber-volume-fractions, whilst also providing dry and wet compaction curves.
This new measurement tool also readily verifies any race-tracking that may occur during impregnation, which
drastically influences flow behavior. The method is successfully validated through testing of glass fiber NCF and
woven fabrics, where permeability results are compared to analytical solutions, while compaction results are
compared to those obtained from a universal testing machine.

1. Introduction volume fraction (Vf). These curves can generally be modeled with a
power law [8,9] or an exponential function [10,11] as follows (where σ
Economical manufacturing routes for fiber-reinforced composites is the applied pressure, A and B are fitting parameters):
such as compression-RTM [1] rely on through-thickness impregnation ( )
σ Vf = A∙exp(B∙Vf ) (1)
strategies. The liquid resin is injected at high pressures into a dry fabric,
which may induce hydrodynamic compaction depending on the injec­ The experimental characterization of unsaturated through-thickness
tion pressure and the fiber-bed compressibility [2,3]. The impregnation permeability (K3unsat) is typically carried out by injecting a test fluid in
behavior and the fiber volume content of the manufactured part are the through-thickness direction. The K3 unsat is then calculated using
governed by two essential parameters: through-thickness permeability Darcy’s law which relates the fluid velocity (v) to the pressure gradient
(K3) and compressibility (dry and wet). Common approaches to char­ (∇p) through the fluid viscosity (μ), Vf and K3unsat as follows:
acterize the compressibility of fibrous reinforcements are displacement
controlled compaction tests [4–6]. The method is applicable to dry or K3unsat
v= − ∙∇p (2)
fully saturated textiles, where the specimens are either immersed in a μ∙(1 − Vf )
bath of the test fluid upon compaction [4,7] or wetted beforehand and Trevino et al. [12] used a universal testing machine to inject a non-
then compressed [5,6]. A considerable difference exists between dry and reactive fluid (DOP oil) into a fiber-bed in through-thickness direction at
wet compressibility of fibrous reinforcements [4–6], where, saturated a constant flow rate, and calculated K3unsat using Darcy’s law. Wu et al.
fiber-beds are more compliant due to lubrication. This facilitates the [13] calculated K3unsat using the same method after improving the
movement of fibers and bundles during the compaction making the previous measurement setup. A central injection was performed from
tested specimen more compressible. Thus, the magnitude of this differ­ the bottom of the cell, and distribution media were added every two
ence is reduced for reinforcements with strong fiber fixation. layers of fiber reinforcement to ensure a 1D flow. In addition, the paper
Compressibility can be quantified by utilizing compaction curves that introduced a new method that combines a 3D flow experiment and
depict the pressure required to compact a fiber-bed to a certain fiber numerical simulation. It requires the prior knowledge of the in-plane

* Corresponding author.
E-mail address: mkabachi@ethz.ch (M.A Kabachi).

https://doi.org/10.1016/j.compositesa.2020.106203
Received 12 June 2020; Received in revised form 5 October 2020; Accepted 14 November 2020
Available online 3 December 2020
1359-835X/© 2021 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
M.A Kabachi et al. Composites Part A 141 (2021) 106203

permeabilities K1 and K2 and compares the simulated and the measured through-thickness permeability (K3sat) was quantified in [19] using a
inlet pressures to back-calculate K3unsat. Later researches used different cylindrical transparent tool and infusing a saturated fiber-bed at
methods to track the flow front position and calculate K3unsat using increasing injection pressures. A decrease of K3sat, due to an increase in
Darcy’s law and applying one of the two following boundary conditions Vf, was reported as a result of the induced hydrodynamic compaction.
depending on the injection strategy: Scholz et al. [20] introduced a continuous method to measure saturated
K3sat by injecting a test fluid into a saturated fiber-bed placed in a cy­
• Constant injection pressure: the difference between the inlet pressure lindrical steel tool mounted on a hydraulic testing machine. The mold
Pin and the pressure at the flow front Pf is constant, resulting in a drop cavity thickness was changed via the testing machine crosshead during
in flow front velocity and a permeability dependent on Vf as the impregnation providing K3sat at different Vf in a single experiment.
following: A comparison between the continuous K3sat results and the values ob­
tained via discrete methods was addressed in [21] via a slightly modified
(1 − Vf )∙μ∙zff 2 setup that uses a universal testing machine to continuously change the
K3unsat = ( ) (3)
2 Pin − Pf ∙t thickness. The difference between the two results is influenced by the
compression speed in the continuous method, it is highly reduced at a
• Constant flow rate injection Q: the flow front velocity is constant, low flow rate and low compression speed [22]. The combination of the
resulting in an increase in inlet pressure over time, and the perme­ parallel but oppositely directed loads on a fiber-bed is modeled in [23]
ability is calculated as follows (A being the specimen cross-section): and could be applied for the Resin Film Infusion process to approximate
the permeability variations during the infusion since the K3unsat is also
K3unsat = (
μ∙Q∙zff (t)
) (4) needed. A K3unsat characterization method that allows for simultaneous
Pin − Pf ∙A tracking of flow front position and total compaction based on US sensors
and linear variable differential transformers (LVDT) was developed in
The tracking of the flow front position zff(t) is essential for unsatu­
[18]. It permits for a continuous K3unsat characterization and considers
rated permeability characterization. Some techniques rely on material-
hydrodynamic compaction but it is limited by the narrow range of Vf
embedded sensors, such as optical fibers or thermistors [8,14]. The
and thicknesses that US sensors can cover.
presence of such sensors inside the fiber-bed creates distortions in the
New measuring techniques are needed to thoroughly characterize
textile’s architecture affecting flow and permeability, thus spatially
K3unsat for the complete range of Vf and a wider span of thicknesses, and
limiting the number of measurement points depending on the number of
fiber architectures (random mat with typically Vf < 40%). The present
sensors [9]. A later non-destructive method uses ultrasound (US) tech­
study is filling this gap by introducing a novel method to characterize
nology [15–18], where US sensors emit and receive sound waves
unsaturated K3unsat considering hydrodynamic compaction. The method
through the fiber-bed in the flow direction during impregnation. Then,
also allows for simultaneous compaction characterization within the
zff(t) is calculated considering the difference in the time of flight (ToF)
same test, using an improved version of the saturated K3unsat measure­
between a dry and a wet fabric, which is defined as the time needed for
ment setup presented in [2,4]. The novel configuration utilizes a PMMA
an emitted sound wave to travel from the emitter to the receptor in a
tube in the sample area that permits visual tracking of zff(t) with a
specific medium. This method allows continuous zff(t) monitoring, but
camera; a well-known technique in the field of unsaturated in-plane
due to signal attenuation, it is limited to measuring only a narrow range
measurements [24–27]. The same optical principle is used to track the
of thicknesses and Vf (above 50%). This limited measuring window can
compaction, which is associated with a pressure sensor allowing for the
be extended to Vf up to 45% by using signal amplifiers [17], which
definition of dry and wet compaction curves and to assess the quality of
further complicates the measurement cell since it is very sensitive to
the experiment via visualization of any race-tracking that may occur
noise.
during the injection.
For Darcy’s law to be used, the permeability and porosity should be
constant, thus K3unsat characterization methods based on through-
2. Methods
thickness injection should be performed at low injection pressures to
avoid hydrodynamic induced compaction [13]. This effect was observed
2.1. Experimental setup
in [12] during a center-gate 3D injection experiment in a mold with a
transparent acrylic window, where local compaction around the injec­
The setup consists of a measurement cell, an injection pump, and a
tion gate was noticed. The influence of such deformation on saturated

Fig. 1. Concurrent through-thickness permeability and compaction characterization setup.

2
M.A Kabachi et al. Composites Part A 141 (2021) 106203

video camera as illustrated in Fig. 1. The measurement cell comprises a


transparent Poly Methyl Methacrylate (PMMA) tube of 123 mm inner
diameter and a wall thickness of 5 mm where the textiles are sandwiched
between two perforated steel plates of 12 mm thickness. The distance
between these plates defines the cavity height (14 mm), which is set
through the height of the PMMA tube. Each plate consists of 95 holes of
8 mm diameter distributed in a hexagonal configuration to ensure low
resistance to fluid flow and even pressure distribution during the in­
jection. The two plates are fitted to the lower and the upper fluid dis­
tribution chambers containing an inlet, outlet and pressure sensor (from
HUBA Type 680.806120105 N) at the bottom. This sensor measures the
inlet pressure Pin(t), which corresponds to the total stress applied to the
fiber bed during impregnation, and is connected to a computer running a
LabView program for the data collection. A video camera (Sony α3 with
90 mm macro lens) is faced towards the PMMA tube to track flow fronts
during impregnation (for K3unsat), as well as the fiber-bed compaction
resulting from hydrodynamic pressure.

2.2. Sample preparation

It is essential to use round samples free of frayed edges to achieve a


tight fit with the PMMA tube’s inner diameter to avoid race-tracking as
shown in [28]. The textiles are precut into squares of 140x140 mm2
using a computer numerically controlled (CNC) textile cutting machine
(Zünd 1600). These square plies are then manually laid up in the same
fiber direction and the same side facing upwards for circular shaping
with a universal testing machine (Zwick). A cylindrical hardened steel
blade is pressed on top of the precut and piled squares to produce cir­
cular stacks of 123 mm diameter, as presented in Fig. 2. In order to avoid
bending of the edges due to the pressure applied by the blade only 10
precut square plies are cut at a time. This procedure allows clean-cut
round samples and avoids losing fibers or bundles when handling sin­
gle layers that can lead to frayed edges. The edges of the first layer are
painted using a color paste to facilitate the visual tracking of the fiber-
bed compaction. Other layers could also be traced to measure any
non-uniform compaction that may occur during impregnation. The Fig. 3. Through-thickness flow, using colored silicon oil, where race tracing
fiber-bed stack is then carefully placed in the measurement cell one or causes peripheral flow around the fiber-bed.
two layers at a time to avoid their bending against the walls assuring a
tight fit.

2.3. Race tracking

Improper sample cutting or inadequate positioning of the layers in


the measurement cell (bending against the cell wall) creates flow
channels between the PMMA wall and the fiber-bed. These channels lead
to an inconsistent flow front and improper impregnation (Fig. 3) and is a

Fig. 4. Inlet pressure development during a through-thickness impregnation of


a woven fabric specimen with Vf0 = 0.40 at a constant flow rate Q = 1.5cm3 /s.

common issue in both in-plane and through-thickness impregnations.


The presented procedure allows for easy detection of even small flow
channels and provides a first qualitative assessment of experiment
validity.
Fig. 2. The hardened steel blade used to cut the samples of 123 mm diameter.

3
M.A Kabachi et al. Composites Part A 141 (2021) 106203

Fig. 5. Illustration of two video frames used for extracting the flow front progression zff(t) and the fiber-bed compaction e(t) during through-thickness impregnation,
with a: before impregnation, b: during impregnation.

2.4. Procedure and data analysis

Injection starts from the bottom at a constant flow rate and stops
when the flow front has reached the upper perforated plate, while a
camera takes a video of the process. The pressure profile over time
(Fig. 4) shows two distinct segments. The red-framed segment is char­
acterized by a nonlinear pressure evolution that can be explained by the
combination of flow in unsaturated porous media, and hydrodynamic
induced compaction. This segment starts as soon as the test fluid reaches
the bottom of the first layer, and ends when the flow front reaches the
top. The red-framed segment is used to determine the K3unsat and the
compaction curve. The blue framed segment corresponds to a saturated
flow, it starts when full impregnation is achieved.
The data analysis includes the conversion of the recorded video into
single frames, wherein each frame shows flow front position zff(t),
sample thickness e(t) at a time t, and the deformation Δe as illustrated in
Fig. 5. These variables are measured using an image analysis Matlab
code based on image segmentation and are used along with the corre­
sponding pressure Pin(t) (red dots in Fig. 4) to calculate permeability
and compaction curves. Fig. 6. Resulting compaction curve (in black) form flow experiment delimited
by dry and wet compaction curves in blue and red respectively.
2.4.1. Compaction curves calculation
To determine compaction curves σ (Vf) the procedure utilizes the following Eq. (8):
exponential model (Eq. (1)), where the constants A and B are calculated σ αm (Vf ) = (1 − αm )∙σ d (Vf ) + αm ∙σ w (Vf ) (8)
using two pressure points P(t) (Fig. 4) and the two corresponding
thicknesses e(t) (Fig. 5) following Eq. (5): By considering two compaction curves σαm1 (Vf) and σ αm2 (Vf) at two
different average impregnation ratios αm1 and αm2, one can deduce dry
P1 ln(P1 /P2 ) and wet compaction curves σ d* (Vf ) and σw* (Vf) respectively following
A= , B= (5)
exp(B⋅Vf1 ) Vf2 − Vf1 Eq. (9):
P1 and P2 are the measured pressures (red dots Fig. 4), Vf1 and Vf2 are αm1 ⋅σαm2 − αm2 ⋅σαm1
calculated considering the thicknesses e(t), the areal weight per layer
σd* =
σ = (1 − αm1 )⋅σd* + αm1 ⋅σw* αm1 (1 − αm2 ) − αm2 (1 − αm1 )
Aw , the number of layers N and the fiber density ρf following Eq. (6): { αm1
σαm2 = (1 − αm2 )⋅σ d* + αm2 ⋅σ w*
⇒{
σαm1 − σd* ⋅(1 − αm1 )
σ w* =
Aw ∙N αm1
Vf = (6)
ρf ∙e(t) (9)

Since the compaction characterization takes place during impreg­ 2.4.2. Unsaturated through-thickness permeability calculation
nation (red segment in Fig. 4), the resulting compaction curve σαm (Vf) The unsaturated through-thickness permeability is calculated based
describes an intermediate state between completely dry σ d (Vf) and fully on Darcy’s law considering a 1D impregnation at a constant flow-rate
saturated σw (Vf) states. In consequence, this curve is bounded by the following Eq. (4), where, zff(t) the averaged flow front position at a
aforementioned curves, as shown in Fig. 6. σαm (Vf) is highly dependent time t considering the compaction as shown in Fig. 5, and Pin(t) is the
on the choice of the two measurement points P1 and P2 taken at two inlet pressure obtained from the pressure sensor. It is assumed in this
different saturation ratios α1 and α2, defined as the ratio between lengths method that the averaged recorded zff(t) around the sample is repre­
of the wet and the dry regions at a timet: sentative of the flow inside because race-tracking is eliminated. This
zff (t) method can consider a high number of measurement points (n) for zff(t)
α(t) = (7) and et e(t), depending on how fast the impregnation occurs, and the
e(t)
number of frames per second the camera can take. Therefore, each
The resulting curve σ αm (Vf ) describes a state with an average satu­ experiment results in n permeability points at n values of Vf since the
ration ratio αm between α1 and α2. This curve will be closer to the dry compaction is considered in the measurement of zff(t).
state if the two measurement points are taken at the beginning of the The compaction in the wet part is not homogenously distributed
impregnation (red points 1 and 2 in Fig. 4), and closer to the wet state if along with the thickness, as shown in the literature [2,29], resulting in a
the two points are taken towards the end (red points 3 and 4 in Fig. 4). gradient of Vf leading to a gradient of K3unsat. Consequently, the
This study assumes therefore that σ αm (Vf) obeys the rule of mixture permeability calculated for each measurement point n represents the

4
M.A Kabachi et al. Composites Part A 141 (2021) 106203

equivalent permeability of the impregnated stack. This is similar to the curves σd* (Vf) and σ w* (Vf) and are compared to dry and wet compaction
equivalent in-plane permeability derived in [30] for a fiber-bed showing curves, σd (Vf) and σ w (Vf) that are obtained from a conventional
a varying local permeability along the flow direction. compaction test with Zwick Materialprüfung 1474 universal material
testing machine as described in [4]. σ αm (Vf) is compared to a reference
3. Experimental compaction curve σref (Vf) calculated following Eq. (8) using the same
saturation ratio αm and the Zwick curves. For this purpose, a stack of 40
3.1. Materials square plies (150x150mm2) of the woven and the NCF glass fibers is
compacted at a speed of 1 mm/min to obtain σ d (Vf) andσw (Vf ). This
Two different textiles are utilized in this work: a twill 2/2 glass fiber speed is defined to match the compaction rate during the impregnation
woven fabric (Hexcel 1202) with an areal weight of 290 g/m2 and a test at the considered flow-rate. The textile layup is inserted between
biaxial ± 45◦ glass fiber non-crimp fabric (NCF) (Saertex X-E-444) with two load steel plates of 135 mm in diameter, where the lower platen is
an areal weight of 444 g/m2. The test fluid used is a low viscosity sili­ fixed, and the upper one is mounted on a load cell attached to a movable
cone oil (XIAMETER® PMX-200, µ = 100 mPa.s at 20 ◦ C), to ensure crosshead. The measurements are controlled via the Zwick Roell testX­
constant viscosity throughout the test. A coloring paste for silicone pert III test software to record displacements and the reaction forces on
(NEUKASIL SN Color Pastes provided by Altropol) is introduced to the the upper load platen using displacement controlled testing. The anal­
oil in a ratio of 20 g/l to readily distinguish oil flow inside the fabrics. ysis consists of five compaction measurements on five different speci­
mens and an average of the curves was evaluated.
The permeability results are compared to analytical solutions
3.2. Test conditions considering the same parameters as Eq. (4) except for the flow front
position zff(t), which is calculated following Eq. (10), with a variable Vf
The experiments are performed on three initial Vf0, {0.40, 0.43, calculated (Eq. (6)) considering the compaction Δe as shown in Fig. 5:
0.45} for the woven and {0.40, 0.43, 0.47} for the NCF, and conducted
Q
on three different samples per Vf0 for both fabrics. The measurements zff (t) = t (10)
(1 − Vf )∙A
are carried out at a constant flow rate of 1.5cm3 /s, which is set through
the injection pump. A cavity height of 14 mm was kept constant and the The cloud of K3unsat results generated by all the experiments is fitted
number of layers was changed to achieve the target Vf0. The video to Gebart’s model [31].
analysis considers 5 measurement points per experiment (5 frames) (√̅̅̅̅̅̅̅̅̅̅̅̅̅̅ )2.5
taken 10 s after the start of impregnation with a time-step of 10 s. Since 16 Va
k3 = √̅̅̅ − 1 ∙r2 (11)
the hydrodynamic compaction is considered, every single experiment 9∙π 6 Vf
results in 5 different K3unsat at 5 different Vf.
The compaction results are validated using the deduced compaction

Fig. 7. Video frames of impregnation of an NCF textile with an initial Vf0 of 0.4 at a constant flow rate of 1.5 cm3 /s at four times (0 s, 10 s, 20 s and 30 s)
demonstrating compaction and flow front progression.

5
M.A Kabachi et al. Composites Part A 141 (2021) 106203

Fig. 8. Pictures of the first impregnated layers of the NCF (Vf0 = 0.55) after an impregnation stopped halfway ordered (from top to bottom).

Where Va = 0.907 is the maximum fiber volume fraction of the hexag­ two different average saturation ratios, αm1 = 20% and αm2 = 50% for
onal fiber arrangement (which is assumed for the test fabrics), r the fiber the woven fabric, and αm1 = 30% and αm2 = 50% for the NCF. The re­
radius. sults are averaged, and compared with dry and wet compaction curves
obtained from a universal testing machine (Zwick) and presented in
4. Results and discussion Figs. 9 and 10 for woven and the NCF respectively along with the
standard deviations at different points.
The video frames in Fig. 7, are excerpts from videos using the newly
developed tool and depict four images from the impregnation of the NCF
textile with an initial Vf0 of 0.4. The robustness of the presented method
is highlighted by the uniformity of the recorded flow front propagation.
To check the validity of the assumptions that the recorded zff rep­
resents the flow inside the fiber-bed and that a tight fit of the sample
eliminates race-tracking, an impregnation experiment of the NCF (Vf0 =
0.55) was stopped halfway by detaching the hose connecting the pump
to the cell. The sample was taken out and the dry layers were removed
until the first partially wet layer was reached, pictures of the next 4
layers are presented in Fig. 8.
Knowing that the thickness of a single layer is approximately 0.3 mm
at the indicated Vf for NCF (calculated eq.6), the pictures presented in
Fig. 8 show that the difference in flow front position between layers 1
and 4 is around 1 mm. This difference is neglected since it is comparable
to the slight fluctuation of the flow front (as shown in Fig. 7 after 10 s or
30 s) which is averaged to provide the zff used for K3unsat calculation.
This figure provides also solid evidence of the absence of race-tracking
since the oil reaches the center and some edges at the same time.

4.1. Compaction

The dry and wet compactions are analyzed during through-thickness Fig. 9. Woven compaction curves obtained at an average α of 20% and 50%
injection using two compaction curves extracted from each flow test at with reference curves from Zwick.

6
M.A Kabachi et al. Composites Part A 141 (2021) 106203

The dry NCF compaction curves, in turn, show a smaller difference at Vf


between 0.50 and 0.54 but remains lower than 20%. The good fit of the
compaction curves with reference results is conclusive evidence in
avoiding race-tracking, as no compaction occurs when the oil bypasses
the specimen, thus resulting in stiffer compaction curves. The afore­
mentioned comparisons validate the novel characterization method and
demonstrate its high potential in characterizing wet and dry compac­
tions during impregnation.

Fig. 10. NCF compaction curves obtained at an average α of 30% and 50%,
with reference curves from Zwick.

The results confirm the hypothesis that the determined compaction


curves represent an intermediate state of impregnation between
completely dry and fully saturated states. The variability (std/σ average )
for both fabrics range from 5% to 13% for Vf between 0.45 and 0.60,
both in dry and wet states. Higher relative errors (around 25%) are
noted at Vf lower than 0.45 and inlet pressure around 104Pa. These
relatively high errors do not affect the robustness of the method since Fig. 11. Woven dry and wet compaction curves deduced from the flow
this range of Vf and injection pressure are too low to be used in pro­ experiment at {(αm1, αm2)}= {(10%, 40%), (20%, 50%), (30%, 60%)} and
ducing engineering parts using NCF or woven fabrics via LCM processes. compared with Zwick.
Measured compaction pressures σαm are compared in Table 1 to
references σref values calculated, based on the rule of mixture, using αm1,
αm2, σd and σ w (from the Zwick) following Eq. (8) for both fabrics at Vf =
0.45 and Vf = 0.60.
Differences between the measured and reference pressures are lower
than 15% for both woven and NCF. This confirms the high consistency of
the obtained compaction curves and supports the assumption that
compaction curves σ αm (Vf) obey the rule of mixture.
The resulted compaction curves from the flow experiments at αm1
and αm2 are used to deduce the dry and fully saturated compaction
curves, respectively σ d* andσ w* , following Eq. (9). To demonstrate that
σ d* andσw* are independent from the choice of αm1 and αm2, three
arbitrary chosen couples {(αm1, αm2)} are utilized, {(10%, 40%), (20%,
50%), (30%, 60%)} for the woven and {(20%, 60%), (30%, 50%), (35%,
45%)} for the NCF. The deduced curves are shown in Figs. 11 and 12 and
compared to σd and σw (from the Zwick) for both woven and NCF.
The deduced compaction curves show good overlapping with the
ones obtained from the Zwick for both fabrics in dry and wet states at the
different couples {(αm1, αm2)}. This confirms that the deduced curves are
independent from the choice of {αm1, αm2}, and validates the assumption
about the application of the rule of mixtures. A small discrepancy is
Fig. 12. NCF dry and wet compaction curves deduced from the flow experi­
noted between σw and σw* with a maximum relative error of 25% for Vf
ment at {(αm1, αm2)}= {(20%, 60%), (30%, 50%), (35%, 45%)} and compared
lower than 0.50 but at low compaction pressures in the range of 20 KPa.
with Zwick.

Table 1
Compaction pressures obtained from flow experiments and reference pressures calculated based on αm1, αm2 σ d and σ w .
Fabric Woven NCF

Pressures σd [bar] σw [bar] σ20%[bar] σ50%[bar] σd [bar] σw [bar] σ30%[bar] σ50%[bar]

Ref Exp Ref Exp Ref Exp Ref Exp

Vf = 0.45 3.2 2.2 3 2.9 2.7 2.4 0.20 0.10 0.17 0.15 0.15 0.13
Vf = 0.60 8.0 3.5 7.1 7.2 5.75 5.9 10.0 6.6 9.0 9.0 8.3 8.0

7
M.A Kabachi et al. Composites Part A 141 (2021) 106203

4.2. Unsaturated permeability technique.

The permeability and Vf results of the three experiments are aver­ 5. Conclusion
aged for each one of the 5 measurement points per Vf0 and presented in
Figs. 13 and 14 along with the standard deviation for the woven and the This study introduces a simple and robust method for the simulta­
NCF, respectively. The permeability results are fitted with Gebart’s neous characterization of compaction (both dry and wet) and unsatu­
model and compared with the analytical model taking into account the rated out-of-plane permeability in one experiment. The measurement
deformation based on Eqs. (5) and (10). setup uses a constant flow rate injection in through-thickness direction,
Figs. 13 and 14 show good agreement between the analytical and the and optically tracks the compaction and flow front position with a
experimental results, wherein the standard deviation for both K3unsat camera, while a pressure sensor evaluates the inlet pressure. The mini­
and Vf is minimal with minor exceptions. The calculated variability is mized standard deviations and successful comparison with the reference
lower than 5% in Vf and lower than 10% in most K3unsat measurements, results demonstrate the consistency of this technique in readily assessing
except the first points at Vf0 = 0.40 and Vf0 = 0.45 for the woven fabric, the compaction behavior of textiles during impregnation. The results
which exhibits a permeability variation of 13% and 25% respectively. show the high potential to replace the current state of the art method­
The increase of the initial fiber volume fraction from 0.406 to 0.436 for ologies where the induced compaction is not considered, that was the
the NCF occurred at 0.25 bar of inlet pressure, resulting in a perme­ case in the last K3 benchmark, where a big scatter is noticeable amongst
ability drop of 25%, from 5.93∙10-12 m2 to 4.45∙10-12 m2. This high­ the participants’ results [32].
lights the importance of considering compaction in K3unsat This technique allows for efficient textile investigations allowing a
characterization. The aforementioned comparisons and curve fittings rapid gathering of data where only one experiment is needed to generate
confirm the robustness and accuracy of the presented measurement reliable information for permeability, dry and wet compactions. A big
advantage of this method is related to the determination of compaction
curves considering different saturation ratios and not just wet and dry.
This makes the characterization of permeability much more realistic for
developing and optimizing real injection processes. The developed
method allows easy detection of race tracking and the observation of the
fluid–structure interactions during the impregnation. It provides both
compressibility and equivalent unsaturated permeability at different Vf
function of time. This allows tailoring of the measurement setup to
closely mimic related manufacturing processes and further optimization
of injection parameters.
The presented method is suitable for various fibrous materials and
covers a very wide range of fiber volume fractions and thicknesses. It
provides an easy robust solution compared to ultrasonic sensing ap­
proaches, which are very sensitive to echo and are applicable only for
relatively thin specimens and a limited range of fabrics and Vf. In
addition, it allows for continuous and non-intrusive measurements,
unlike the embedded-sensors approaches (thermistors or optical fibers),
which provide spatially discrete measurements depending on the posi­
tion of the sensors and their number that could severely affect the flow.
The method will be even more robust by using more than one camera
pointed at different angles of the measurement cell. This would provide
Fig. 13. Experimental and analytical unsaturated K3unsat of the woven fabric
more information about the flow front all around specimens and helps to
fitted with Gebart’s model with r = 10.5 µm. increase the precision of the measurements.

CRediT authorship contribution statement

M.A Kabachi: Conceptualization, Methodology, Validation, Formal


analysis, Investigation, Data curation, Writing - original draft, Writing -
review & editing. L. Stettler: Investigation, Data curation. S. Arreguin:
Writing - review & editing, Supervision. P. Ermanni: Writing - review &
editing, Supervision, Funding acquisition.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Acknowledgment

The authors would like to thank the Swiss National Scientific


Foundation (SNF) for funding the research (reference 2-77114-16) that
had led to these results.

Fig. 14. Experimental and analytical unsaturated K3unsat results of the NCF
fabric fitted with Gebart’s model with r = 11.5 µm.

8
M.A Kabachi et al. Composites Part A 141 (2021) 106203

References [16] Thomas S, Bongiovanni C, Nutt SR. In situ estimation of through-thickness resin
flow using ultrasound. Compos Sci Technol 2008;68:3093–8.
[17] Konstantopoulos S, Grössing H, et al. Determination of the Unsaturated Through-
[1] Moon-Kwang U, Lee WI. A study on the mold filling process in resin transfer
Thickness Permeability of Fibrous Preforms Based on Flow Front Detection by
molding. Polym Eng Sci 1991;31:NO. 1.
Ultrasound. Polym Compos 2018.
[2] Klunker F, Ermanni P, et al. Fiber deformation as a result of fluid injection:
[18] Willenbacher B, May D, Mitschang P. Metrological determination of
modeling and validation in the case of saturated permeability measurements in
inhomogeneous hydrodynamic compaction during unsaturated out-of-plane
through-thickness direction. J Compos Mater 2015;49(9):1091–105.
permeability measurement of technical textiles. Adv Manuf Polym Compos Sci
[3] Willenbacher B, Kabachi MA, et al. Flow induced sample deformations in out-of-
2019;5(2):51–4.
plane permeability measurement. In: 18th European Conference on Composite
[19] Endruweit A, Luthy T, Ermanni P. Investigation of the influence of textile
Materials (ECCM 2018), Athens, Greece.
compression on the out-of-plane permeability of a bidirectional glass fiber fabric.
[4] Kabachi MA, Ermanni P, et al. Experimental study on the influence of cyclic
Polym Compos, August 2002, vol. 23, no. 4.
compaction on the fiber-bed permeability, quasi-static and dynamic compaction
[20] Scholz S, Gillespie Jr JW, Heider D. Measurement of transverse permeability using
responses. Compos A 2019;125:105559.
gaseous and liquid flow. Compos A 2007;38:2034–40.
[5] Kelly PA, Umer R, Bickerton S. Viscoelastic response of dry and wet fibrous
[21] Ouagne P, Bréard J. Continuous transverse permeability of fibrous media. Compos
materials during infusion processes. Compos A 2006;37:868–73.
A 2010;41:22–8.
[6] Comas-Cardona S, Le Grognec P, Binetruy C, Krawczak P. Unidirectional
[22] Ouagne P, et al. Continuous measurement of fiber reinforcement permeability in
compression of fibre reinforcements. Part 1: A non-linear elastic-plastic behaviour.
the thickness direction: Experimental technique and validation. Compos B 2013;
Compos Sci Technol 2007;67(3):507–14.
45:609–18.
[7] Danzi M, Klunker F, Ermanni P. Experimental validation of through-thickness resin
[23] Ouahbi T, et al. Modelling of hydro-mechanical coupling in infusion processes.
flow model in the consolidation of saturated porous media. J Compos Mater 2017;
Compos A 2007;38:1646–54.
51(17):2467–75.
[24] Vernet N, et al. Experimental determination of the permeability of engineering
[8] Ballata WO, Walsh S, Advani S. Determination of the Transverse Permeability of a
textiles: Benchmark II. Compos A 2014;61:172–84.
Fiber Preform. J Reinf Plast Compos 1999;18(16).
[25] David May et al. In-plane permeability characterization of engineering textiles
[9] Sirkis JS, Dasgupta A. The Role of Local Interaction Mechanics in Fiber Optic Smart
based on radial flow experiments: A benchmark exercise. Composites. Part A, vol.
Structures. J Intell Mater Syst Struct 1993;4:260.
121, pp. 100–114, Oxford: Elsevier, 2019.
[10] Somashekar AA, Bickerton S, Bhattacharyya D. Modeling the viscoelastic stress
[26] Swery E, Comas-Cardona S, et al. Efficient experimental characterisation of the
relaxation of glass fibre reinforcements under constant compaction strain during
permeability of fibrous textiles. J Compos Mater 2016;50(28):4023–38.
composites manufacturing. Compos A Appl Sci Manuf 2012;43(7):1044–52.
[27] Gourichon B, Binetruy C, et al. Experimental investigation of high fibre tow count
[11] Ahn SH, Lee WI, Springer GS. Measurement of the Three-Dimensional Permeability
fabric unsaturation during RTM. Compos Sci Technol 2006;66:976–82.
of Fiber Preforms Using Embedded Fiber Optic Sensors. J Compos Mater 1995;29:
[28] Lawrence J, Barr J, Karmakar R, et al. Characterization of preform permeability in
714.
the presence of race tracking. Compos A 2004;35:1393–405.
[12] Trevino L, James Lee L. Analysis of resin injection molding in molds with preplaced
[29] Michaud V, Manson J-AE. Impregnation of Compressible Fiber Mats with a
fiber mats. I: permeability and compressibility measurements. Polymer
Thermoplastic Resin. Part I: Theory. J Compos Mater 13/2001.;35:No.
Composites, February 1991, Vol. 12, No. 1.
[30] Fratta C, Klunker F, Trochu F, Ermanni P. Characterization of textile permeability
[13] Wu C-H, James Wang T, James Lee L. Trans-plane fluid permeability measurement
as a function of fiber volume content with a single unidirectional injection
and its applications in liquid composite molding. Polym Compos, August 1994,
experiment. Compos A 2015;77:238–247239.
Vol. 15, No. 4.
[31] Gebart B. Permeability of unidirectional reinforcements for RTM. J Compos Mater
[14] Weitzenbock JR, Shenoi RA, Wilson PA. Determination of three-dimensional
1992;26(8):1100–33.
permeability of fiber preforms by the inverse parameter estimation technique.
[32] May D, Aktas A, Yong A. International benchmark exercises on textile permeability
Compos Part A: Appl Sci Manufact 1998;29:159.
and compressibility characterization. ECCM18 - 18th European Conference on
[15] Stöven T, Weyrauch F, Mitschang P, Neitzel M, Stöoven T, Weyrauch F, et al.
Composite Materials Athens, Greece, 24-28th June 2018.
Compos A 2003;34:475–80.

You might also like