Vous êtes sur la page 1sur 12

Original Article

Journal of Reinforced Plastics


and Composites
2016, Vol. 35(24) 18021813
! The Author(s) 2016
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/0731684416669249
jrp.sagepub.com

Capturing the influence of geometric


variations on permeability using a
numerical permeability prediction tool
Elinor E Swery1, Tom Allen1 and Piaras Kelly2

Abstract
An automated tool has been developed for generation of permeability predictions for multi-layered unit cells utilising
textile modelling techniques. This tool has been used to predict the permeability tensor of a woven textile. Single-layer
predictions were carried out and the predicted permeabilities obtained were in close agreement to the permeability
behaviour captured experimentally. The tool was used to capture the effects of textile variability on its permeability,
isolating the influence of individual parameters. A complete textile sample was also analysed, predicting its permeability map. The concept of estimating the permeability of a textile with variability using an average single unit cell was
explored. The prediction tool was also used to study the effect of preform structure on its permeability, including
consideration of the number of layers, ply shift and applied compaction.

Keywords
Permeability, textile modelling, liquid composite moulding

Introduction
Fibre-reinforced polymer composite (FRPC) materials
are used in a large number of industrial applications.
FRPCs consist of two or more distinct materials
(generally bre reinforcement and polymer matrix),
forming a material with more desirable properties
compared to the constituent materials. For applications
where parts are mass produced and high levels of
dimensional accuracy and repeatability are required,
liquid composite moulding (LCM) processes are the
preferred manufacturing method.1 In LCM processes,
resin is injected into a mould containing the dry brous
reinforcement and is left to cure before the nal part
can be removed from the mould.
The use of LCM simulation as process design tools is
increasing in industry. Such simulations aim to accurately predict ll time, ow front advancement and dry
spot formation, ultimately enabling the production of
complex high quality parts using the most ecient conditions.2,3 In these simulations, the resin ow during the
LCM manufacturing process is commonly modelled
using Darcys law.4 The relation is given in equation
(1), where q is the volume averaged Darcy velocity; ,

the uid viscosity; P, the uid pressure and K, the permeability tensor. These simulations therefore require
knowledge of the reinforcing materials permeability
characteristics.
q

1
KrP


Permeability, K, is a measure of the ability of a


reinforcement material to transmit uids. As the permeability behaviour of reinforcing textiles is a strong
function of the textiles complex architecture,5,6 which
is extremely dicult to model analytically, many

1
Department of Mechanical Engineering, Centre for Advanced
Composite Materials, University of Auckland, Auckland, New Zealand
2
Department of Engineering Science, University of Auckland, Auckland,
New Zealand

Corresponding author:
Elinor E Swery, Department of Mechanical Engineering, Centre for
Advanced Composite Materials, University of Auckland, Private Bag
92019, Auckland 1142, New Zealand.
Email: elinor.swery@auckland.ac.nz

Swery et al.

1803

Automated tool (executed within Matlab)

Textile
Image

Matlab

TexGen

TexGen

TexGen

Abaqus

Matlab

HyperMesh

ANSYS CFX

Image
Analysis

Textile
Model

Preform
Model

Dry Fibre
Mesh

Compaction
Simulation

Create Voxel
Mesh

Prepare
Mesh

Flow
Simulations

.csv

.tg3

.tg3

.inp

.vrml

.inp

.cas

.wbpj

Permeability

Figure 1. Automated permeability prediction tool steps.

researchers are relying on measurements obtained from


experiments. For this, many data points are required to
capture the inuence of manufacturing and handling
induced geometric variability, varying compaction
levels, and the preform structure. This is time and
labour intensive and the results can vary signicantly
due to a lack of standardised testing methods and
errors involved in the handling process.7,8
Another method used to obtain the permeability
characteristics of reinforcing textiles is through
numerical simulation approaches. Textile modelling
techniques can be used to generate models of a representative unit cell of the reinforcement. By imposing a
pressure drop across the unit cell and solving the governing ow equations, the permeability properties may
be determined.9 A number of methods are currently
available that enable the implementation of textile
modelling techniques. The most common are the
WiseTex suite, developed at KU Leuven10 and
TexGen, developed at the University of Nottingham.11
Numerous works have presented permeability predictions obtained using textile modelling techniques919; however, few of these theoretical
researches include the eect that textile geometric variability has on the permeability. It was previously shown
that technical textiles do exhibit a signicant amount of
variation2023 and that this variation does inuence the
textiles permeability.24,25 Hence, this paper examines
the inuence that geometric variations have on the textiles permeability. This is an aspect that is otherwise
very dicult to control and study experimentally.
An automated tool that has been developed to predict the permeability of reinforcing textiles is presented.
A signicant emphasis has been placed on ensuring that
the process is fully automated, fast and ecient, requiring minimal input data. The various permeability predictions were carried out on an 800 gsm balanced plain
woven E-glass textile (supplied by Gurit Asia Pacic)
compacted to three target bre volume fractions (Vf),
25%, 34% and 47%. Both single-layer and multi-layer
preform structures (comprising 2, 4, 8 and 10 layers)
were analysed, providing further insight into the associated nesting behaviour of the textile tows during
compaction.

Automated permeability prediction tool


An automated tool used to generate the permeability
predictions has been developed utilising textile modelling techniques. Figure 1 outlines the steps involved in
this prediction process. These steps are performed directly from the MATLAB environment without any user
intervention. When a step is completed using an external software package, appropriate scripts are generated
and executed from within MATLAB. For example, a
python script is created to execute the compaction
simulations in Abaqus.26 A general script is modied
for the specic simulation so that the correct parameters (compaction distance, boundary condition
regions, material properties, etc.) are used. Such an
automated tool enables the ecient analysis of a large
number of cases.
Data regarding the textiles architecture has been
obtained using automated image analysis techniques,
as previously presented by Swery et al.20 Images of the
textile were obtained using a standard oce scanner
and the in-plane geometry of the textile
determined from automatically identifying the edges
of the tows. Through-thickness data about the tows
(such as their thickness and cross-sectional shape)
were obtained from microscopy images. These data
sets were used to generate the unit cell geometries in
TexGen, an open source software licensed under the
General Public License, developed by the
University of Nottingham.11 Unit cell models were
used, whereby the repeating structure of the textile
is represented using the smallest possible model,27
substantially reducing the associated computational
expense.28
An approach used to model textiles and their permeability, previously presented by Takano et al.29 and
Staan Lundstrom30. Note that this assumption implies
that there is no intra-yarn ow. This approximation to
actual events is valid provided that the fabric is compacted to volume fractions which are not too high, typically below about 50%; in that case, the dierence in
permeability predicted by models incorporating permeable tows and those involving solid tows will be small
relative to the permeability changes seen with changes

1804

Journal of Reinforced Plastics and Composites 35(24)

Figure 2. Textile models. (a) single-layer plain woven, (b) four-layers stacked directly on top of each other, (c) four-layers with
maximum ply shift and (d) four-layers with random ply shift.

Figure 3. Four-layer textile model with maximum ply shift (Figure 2(c)), compacted to a range of thicknesses. Resulting Vf of (a) 25%,
(b) 34% and (c) 47%.

in volume fraction.31 The cross-section shape is dened


as the smallest ellipse that encompasses all the bres
within the tows. To model multi-layer preform stacks,
the single-layer models are combined using the stacking
sequence desired. An in-plane shift, i.e. a displacement
of a ply in-plane, can also be applied, thereby controlling the resulting level of nesting exhibited. Figure 2
illustrates a range of textile models that have been created using this method.
Compaction simulations are applied to the textile
models using the Abaqus explicit solver, reecting the
change of geometry due to the compaction (mould closing) step of the LCM process (see literatures3235 for
similar works in this area, which outline the general
approach). The textile mesh is compacted between
two rigid platens to achieve a target cavity thickness.
Transverse tow stiness and Poisson ratios have been
obtained from compaction testing of single tows. These
values have been applied to the tows in the compaction
models. Numerical simulation of a single tow has been
undertaken and compared with experimentally determined deformation to validate this approach. For the
material studied, an elastic modulus of 117 MPa and a
Poissons ratio of 0.31 were used. Figure 3 shows examples of the four-layer preform stack shown in
Figure 2(b) compacted to dierent cavity thicknesses.
An algorithm to automatically generate a voxel mesh
representing the volume of resin surrounding the compacted mesh geometry, and compatible with ANSYS
CFX uid solver, is then applied; the algorithm
accounts for cases where the compacted mesh (which
forms the boundary of the uid region) contains small
element intersections and automatically cleans the mesh

to delete any oating elements (dened as voxels that


are not connected to the main mesh, typically less than
0.2% of total mesh volume). This enables the ecient
execution of uid ow simulations. The boundary
regions are also dened, where the appropriate elements for each region are located. This is achieved by
intersecting two-dimensional planes at the corresponding locations with the mesh.
Flow simulations are carried out in ANSYS CFX,
with a pressure gradient imposed between the inlet and
outlet boundaries. A pressure gradient of 200 kPa/m
was used in all cases (with the outlet pressure set to
zero), in which case the inlet pressure for the average
unit cell was 2500 Pa (as in the experiments); pressure
changes were studied and were seen to have a steady
and predictable eect on the average velocity for such
creeping ow (Re < 1), it is accepted that there should
be no dierence on calculated permeability. Non-slip
wall conditions were imposed on the top and bottom
surfaces of the unit cell as well as the boundaries
between the uid domain and the bre tows.
Translational periodic boundaries were used for the
two sides, to reect the periodic structure of the textile.36 These boundary conditions (as used to determine
permeability in the x direction) are illustrated on a
single-layer model in Figure 4. The uid parameters
used reected the properties of typical thermoset
resins used in LCM processes. A viscosity value of
kg
 0.1048 Pa.s and density  870 mm
2 were used in
the simulations, to match that of the oil used in the
experiments (a wide range of viscosity and density
values was tested; as expected, similar results were
_
obtained in all cases). The resulting mass ow rate, m,

Swery et al.

1805

was used to compute the unit cell meso-scale permeability through the application of Darcys law5 as shown in
equation (2).
K

_
mL
AP

In order to compute the permeabilities of the unit


cell in the dierent directions (e.g. weft and warp), the
boundaries were switched; however, the same mesh
used. This is shown in Figure 5 where the same mesh
(generated from the compacted bre mesh shown in
Figure 3(b)) has been used to predict the full threedimensional (3D) permeability tensor. Velocity contour
plots are presented, highlighting the ow channels in
the preform structure.
Figure 6 presents the associated processing times to
complete each step on the single-layer unit cell model

presented in Figure 2(a). These times were achieved by


executing the process on a Windows 7, 64 bit computer,
with 8GB RAM and an Intel CoreTM i5-2500 CPU at
3.30 GHz. As can be seen, the entire process takes
4.7 min, making the permeability prediction tool
(including scanning and image analysis) much more
time-ecient option than carrying out experiments.
The longest components of the process are, as expected,
the compaction and ow simulations.
All permeability predictions were veried by comparing with data obtained experimentally (see Swery
et al.37 for details of experimental procedures and full
results). In-plane data is obtained using an existing inplane, radial injection permeability rig where the ow
front is tracked visually.38,39 The through-thickness
data are obtained from tests conducted on an existing
rig, where the testing uid is injected through the
sample at a given pressure gradient and the resulting
mass ow rate, which is used to compute the permeability, is measured.38

Single-layer permeability predictions


INLET:
Opening Boundary
P(inj)

LEFT:
Translational Periodic
Fluid-fluid

The complete permeability prediction tool has been


used to generate a range of permeability predictions
for the 800 gsm plain woven textile. This section focuses
on predictions made on single-layer textile samples,
compacted to a range of bre volume fractions.

Average unit cell


OUTLET:
Opening Boundary
P(atm)

RIGHT:
Translational Periodic
Fluid-fluid
TOP, BOTTOM, TOWS:
Non-slip walls

Figure 4. Boundary conditions for flow simulations in the x


direction.

Six 250  250 mm textile samples from the same roll


were analysed, capturing approximately 2000 unit
cells (dened as a textile model with two weft and
two warp tows). The geometric parameters obtained
were used to create a global average, single-layer, unit
cell model to represent the textile. This model was used
within the complete permeability prediction tool to
obtain the in-plane permeability predictions at dierent

Figure 5. Velocity contour plots of flow simulations carried out on the voxel mesh generated from the compacted fibre mesh shown
in Figure 3(b). Flow simulations in the (a) x, (b) y and (c) z directions.

1806

Journal of Reinforced Plastics and Composites 35(24)

compaction levels. The permeability results in each of


the x and y directions, as well as the anisotropy ratios,
are shown in Figure 7 for each of the target bre
volume fractions (achieved through dierent levels of
compaction). The lowest Vf, 25%, is of the uncompacted textile model. The predicted values are compared to average experimental values obtained for
single-layer textile samples. The experimental errors
bars represent one standard deviation.
As can be seen in Figure 7, the permeability predictions accurately capture the decreasing permeability
trend with increasing Vf (increasing level of

2.77
38.8

132
67.2

compaction). The predicted permeability values are


very close to all the experimental results in both the x
and y directions.
The ratios between the predicted and experimental
permeability values in the x direction range between
0.78 and 1.28. In the y direction, they range between
0.98 and 1.26. In both cases, the predictions at the highest Vf are lower than the experimental value. This can
likely be attributed to the manner in which the compaction simulations capture the change in textile geometry.
At the higher compaction levels, the tows would spread
more in the compaction simulations (thereby reducing
the gaps and resulting in a decreased permeability),
rather than undergo a reduction in the tow porosity
level. This is likely to have caused the greater decrease
in permeability that was not observed experimentally.
The in-plane anisotropy ratios, presented in
Figure 7(c) also show close agreement between the predicted and experimental values. All predicted in-plane
anisotropy ratios are within one standard deviation of
the experimental mean which shows the prediction
tools ability to accurately capture the permeability
characteristics of the textile.

Geometric variations
13.3
28.3
Textile Model
Dry Fibre Mesh
Compaction Simulation
Create Voxel Mesh
Prepare Mesh
Flow Simulations
Figure 6. Simulation times (s) for compacted, single-layer unit
cell.

(b)
Simulations
Experiments

10 9

10 10

10 8

(c)
Simulations
Experiments

10 9

10 10

3.5
3

Anisotropy Ratio

10 8

Permeability [m2]

Permeability [m2]

(a)

The inuence of the textiles geometric variations on the


permeability has been assessed. A range of dierent
unit cells were generated using the average and spread
of the geometric parameters captured. Within a given
sample, the warp and weft tow widths, wwarp and wweft
and tow gaps, swarp and sweft were altered separately,
maintaining the other geometric parameters at their
identied mean values. Each parameter was tested at
its mean value, as well as 1, 2 and 3 standard deviations (where possible). These parameters are illustrated in Figure 8.

Simulations
Experiments

2.5
2
1.5
1
0.5

10 11

25
34
47
Target Fibre Volume Fraction [%]

10 11

25
34
47
Target Fibre Volume Fraction [%]

25
34
47
Target Fibre Volume Fraction [%]

Figure 7. Single-layer permeability results. Average unit cell results compared with experimental results. (a) Kxx, (b) Kyy and (c)
anisotropy ratio.

Swery et al.

1807

These unit cells were examined at each of the dierent bre volume fractions. The achieved Vf does not
remain constant for all unit cells with geometric variations. Instead, the unit cells were compacted to a constant target cavity thickness, matching those employed
experimentally to achieve the target Vf of interest.
Tow width. The inuence that the tow widths have on the
predicted permeability is illustrated in Figures 9 and 10.
Both the warp (x) and weft (y) tow widths were varied
and the resulting permeabilities in the x and y directions
are shown for the dierent tested values of Vf. The
average experimental results are also shown (dashed
line) for comparison purposes. It is clear that changing

the tow widths has minimal eect on the resulting permeability in both directions. This is as expected, as
altering the tow widths does not change the unit cells
main ow channels, hence the permeability is
unaected.
Gap width. The unit cell gap width has a signicant
inuence on the resulting permeability. This is as
expected, as the gap between the tows determines the
size of the ow channel, directly aecting the permeability. This inuence is illustrated in Figures 11 and 12.
As shown in these gures, increasing the gap width
between the warp tows, signicantly increases the permeability in the x direction while only slightly increasing the permeability in the y direction. The former
occurs as the equivalent diameter of the ow channels
increase. In the latter case, the overall Vf of the unit cell
decreases, which contributes to the slight increase in
permeability. Similarly, an increase in the weft gap
width signicantly increases the permeability in the y
direction while only slightly increasing the permeability
in the x direction.

Variability in complete textile sample


The permeability prediction tool was also used to generate complete permeability maps of the textile studied.
A 250  250 mm sample of the textile was scanned and
analysed to determine its geometry. A total of 420 small
unit cells were then generated based on this geometry.
The permeability of each was determined using the permeability prediction tool. These results were used to

Permeability [m2]

(a)

(b)

Permeability [m2]

Figure 8. Plain woven textile architecture with dimensions of


interest.

10

10

10

10

10

10

10

10

V = 25%

Vf = 25%

Vf = 34%

V = 34%

11

10

Vf = 47%
4

4.5

5.5

Warp (x) Tow Width [mm]

Vf = 47%

11

10

4.5

5.5

Warp (x) Tow Width [mm]

Figure 9. (a) Kxx and (b) Kyy of unit cells with varying warp (x) tow widths. Simulation results along with average experimental results
(dashed line). As noted in the text, these variations induce small variations to the unit-cell volume fractions (which are nominally
constant).

Permeability [m2]

(a)

Journal of Reinforced Plastics and Composites 35(24)

10

10

10

(b)

Permeability [m2]

1808

10

10

10

10

10

Vf = 25%

Vf = 25%

V = 34%

V = 34%
f

10

V = 47%

Vf = 47%

11

4.5

5.5

10

6.5

11

4.5

5.5

6.5

Weft (y) Tow Width [mm]

Weft (y) Tow Width [mm]

Figure 10. (a) Kxx and (b) Kyy of unit cells with varying weft (y) tow widths. Simulation results along with average experimental
results (dashed line). As noted in the text, these variations induce small variations to the unit-cell volume fractions (which are
nominally constant).

10

10

10

(b)

Permeability [m2]

Permeability [m2]

(a)

10

10

10

10

10

V = 25%

V = 25%

Vf = 34%

Vf = 34%

10

Vf = 47%

11

0.5

1.5

2.5

Vf = 47%

11

Warp (x) Gap Width [mm]

10

0.5

1.5

2.5

Warp (x) Gap Width [mm]

Figure 11. (a) Kxx and (b) Kyy of unit cells with varying warp (x) tow gap widths. Simulation results along with average experimental
results (dashed line). As noted in the text, these variations induce small variations to the unit-cell volume fractions (which are
nominally constant).

generate permeability maps of the single-layer textile,


highlighting the inuence that the variability within the
sample has on its permeability. An example of this is
shown in Figure 13, where the permeability in the y
direction for the sample compacted to a target Vf of
47% is shown. The textiles tow edges have been superimposed onto the contour plot for illustrative purposes.
This contour plot shows the signicant variation in
permeability in the y direction. It can be seen that the
locations of higher permeability (shown in red) correspond to locations where the vertical gaps between the
reinforcement tows are larger. This is as expected:

larger gaps provide bigger ow channels for the uid,


thereby increasing the textiles permeability.
Additional contour plots are shown in Figure 14.
The permeabilities in the x and y directions are shown
along with the unit cell angle (which is equivalent to the
permeability ellipse angle of rotation). The predicted Vf
is also shown and has been estimated using equation
(3), where VM is the volume of the matrix mesh and VU
is the total unit cell volume. To account for the porosity
of the tows (which are modelled as solid volumes here),
this is multiplied by their approximate bre volume
fraction fraction40; Vf is taken to be 0.5 here, obtained

Swery et al.

1809

10

(b) 108

Permeability [m2]

Permeability [m2]

(a) 108

10

10

10

10

10

V = 25%

Vf = 25%

Vf = 34%

Vf = 34%

11

10

V = 47%

Vf = 47%
0

0.5

1.5

10

11

0.5

1.5

Weft (y) Gap Width [mm]

Weft (y) Gap Width [mm]

Figure 12. (a) Kxx and (b) Kyy of unit cells with varying weft (y) tow gap widths. Simulation results along with average experimental
results (dashed line). As noted in the text, these variations induce small variations to the unit-cell volume fractions (which are
nominally constant).

x 10
5

10

2500

2
1000

yy

3
1500

Unit Cell K [m2]

y Location [px]

4
2000

500
500

1000 1500 2000


x Location [px]

2500

Figure 13. Kyy contour plot of textile at a target Vf of 47% with


superimposed tows. (This was created using 420 individual permeability values, each value being computed from a single unit cell
whose geometry was derived from the geometry of the large
sample at that location.)

from tow mass and volume data.40 The volume fraction


is evaluated according to equation (3). Note that the
variation in volume fraction clearly evident in
Figure 14(d) is not dependent, except to a small
degree, on the accuracy with which the volume fraction
is determined through equation (3).
Vf vf

VU  VM
VU

These contour plots highlight the importance of


incorporating geometric variability when predicting

permeability for the use in process simulations. The


permeability is seen to vary widely, with dierences of
well over 500% seen in any single sample, and this is
attributed to the inherent geometric variability of the
textile. As expected, regions of higher Vf correspond to
reduced permeabilities in both the x and y directions. A
defect in the textile is identied in the top left corner.
Here, the fabric has been deformed slightly and this is
shown by the tows outline in Figure 13. The unit cell
angle (Figure 14(c)) is noticeably dierent to the rest of
the material, and the resulting permeabilities in the x
and y directions are higher than anticipated.
The 420 individual permeability predictions made
from the large textile sample are also plotted in
Figure 15, showing the large range of results obtained.
Additionally, the in-plane permeability data obtained
from experiments is shown by the dotted lines. The
permeability tests were conducted on the same textile
sample right after it was scanned for the permeability
predictions. The sample was scanned when already
placed on the experimental facility, hence ensuring
that the textile did not deform any further due to handling. Figure 15 shows that the experimental values are
within the range predicted numerically.
In addition to the 420 unit cells that were analysed,
an average unit cell (created from the average geometric
parameters of the corresponding textile sample) was
analysed. The in-plane permeability results obtained
from the analysis of this are also shown in Figure 15.
These results are very similar to the average permeability values of all the individual unit cells. It can be seen
how, by carrying out the permeability prediction process on a single average unit cell, a representative value
may be obtained. However, in order to obtain a

1810

Journal of Reinforced Plastics and Composites 35(24)

(a)

x 10
5

10

(b)

x 10
5
2500

2500

2
1000

500

1000 1500 2000


x Location [px]

3
1500
2

1000

500

2000

500
500

2500

(c)

1000 1500 2000


x Location [px]

2500

(d)

2500

yy

3
1500

y Location [px]

Unit Cell Kxx [m ]

2000

Unit Cell K [m2]

4
y Location [px]

10

0.52
2500

1500
2
1000

2000
0.48
1500
0.46

Unit Cell Vf

y Location [px]

y Location [px]

2000

Unit Cell Angle [ ]

0.5
4

1000
0.44

1
500

500
500

1000 1500 2000


x Location [px]

500

2500

1000 1500 2000


x Location [px]

0.42

2500

Figure 14. Textile permeability maps for textile compacted to a target Vf of 47%. (a) Kxx, (b) Kyy, (c) unit cell angle and (d) achieved Vf.

complete understanding of the textiles permeability,


numerous predictions have to be completed.

10

Simulations Kxx
Simulations K

yy

The complete 3D permeability tensor was predicted for


a range of multi-layer preforms using the average textile
unit cell (created from the geometric parameters previously captured in Swery et al.20). These preforms were
created using 1, 2, 4, 8 and 10 layers, which enabled the
inuence of number of layers to be assessed. Dierent
in-plane ply shifts (as previously displayed in Figure 2)
were applied: no shift, whereby the layers are stacked
directly on top of each other, a maximum shift,
enabling maximum nesting to occur, and a random
shift. The Kxx, Kyy and Kzz of each of these preforms
were then predicted for three target Vfs.
Experimental results are used to validate the predicted results. It is important to note that no special
attention was given to the stacking method during the
experiments; the orientation of the layers was maintained the same but the in-plane shift between the
layers was not controlled.

Average Unit Cell K

xx

Average Unit Cell K

2
Permeability [m ]

Multi-layer permeability

yy

Average of all K

xx

Average of all K

yy

10

10

0.4

0.45

0.5

0.55

Fibre Volume Fraction

Figure 15. Comparison of all Kxx and Kyy values with experimental values (dashed lines) for a single large textile sample.

In-plane
Kxx is displayed in Figure 16. The predicted permeability decreases for each of the preforms examined with

Swery et al.

10

10

10

Directly on top
Random ply shift
Maximum ply shift
Experiments

11

(c)

10

Permeability [m2]

10

(b)

10

Permeability [m ]

Permeability [m ]

(a)

1811

10

10

10

Number of Layers

Directly on top
Random ply shift
Maximum ply shift
Experiments
10

10

10

Directly on top
Random ply shift
Maximum ply shift
Experiments

11

10

10

10

11

10

10

Number of Layers

10

Number of Layers

Figure 16. Kxx, varying number of layers and stacking methods. Compared with experiments. Compacted to a target volume
fraction of (a) 25%, (b) 34% and (c) 47%.

10

10

10

10

10

10

10

11

Directly on top
Random ply shift
Maximum ply shift
Experiments

12

13

10

10

10

Number of Layers

10

(c)

10

Directly on top
Random ply shift
Maximum ply shift
Experiments

10
2

2
Permeability [m ]

10

(b)

Permeability [m ]

10

Permeability [m2]

(a)

10

10

11

Directly on top
Random ply shift
Maximum ply shift
Experiments

12

13

10

10

11

10

12

10

13

10

Number of Layers

10

10

Number of Layers

Figure 17. Kzz, varying number of layers and stacking methods. Compared with experiments. Compacted to a target volume fraction
of (a) 25%, (b) 34% and (c) 47%.

increasing Vf. It can be seen that, in general, the predicted permeabilities appear to asymptote with increasing number of layers. This is as expected, as the
contribution of the boundary eects at the compression
platens on the preform lessens.
The eect of layer shift is also evident by comparing
the predicted permeability results of the preforms made
using the three dierent shifting methods, i.e. perfectly
aligned tows, fully nested and an intermediate ply-shift.
Figure 16 shows that the permeability at the lower Vf is
heavily inuenced by the stacking method. The
preforms that were created with maximum ply shift
generally exhibit the lowest permeability as the alternating placements of the tows obstruct the ow channels (as shown in Figure 3). In contrast to this, the
preforms that were created by stacking the layers directly on top of each other generally exhibit the highest
permeabilities, as larger ow channels are created.
As expected, the random ply shift permeability results
generally lay between the limits created from the other
two preform structures.
The experimental results are also shown on
these plots. These expose the inherent variability in

permeability: in Figure 16(a), for example, the


experimental results for permeability increase with
number of layers initially, but then decrease. This
appears to be a result of variability in layer-shifting, a
combination of the monotonic increase seen with perfectly aligned layers and the slight monotonic decrease
seen with perfectly nested layers. The experimental
results inherent variability might explain why, in
some cases, the experimental results are not within
the range obtained using the prediction tool. Despite
this, in general, the experimental results follow similar
trends as observed in the predicted results and their
values generally lie within the predicted spread.

Through-thickness
To compute the through-thickness permeability, the
pressure gradient is applied along the z direction and
two sets of translational periodic pairs are dened in
both of the in-plane directions (see Figure 4). The predicted through-thickness permeability results, Kzz, are
provided in Figure 17. As can be seen, the permeability
of the preforms created by stacking layers directly on

1812
top of each other does not decrease signicantly with
increasing Vf. In the case of these preforms, large,
prominent ow channels are present as the gaps
between the tows align in the thickness direction. As
compaction is applied to the stack, the size of these
gaps reduces slightly due to the tows Poissons ratio
eect, but not signicantly, hence the permeability is
only slightly aected.
In contrast to this, the permeability of the preform
created with maximum ply shift does decrease noticeability with increasing Vf (in particular for the highest
Vf tested). This preform has the most signicant interply nesting present. With increasing compaction
applied to the sample the gaps between the tows are
obstructed by the alternating tows, reducing the ability
of the uid to travel along the original ow path.
Interestingly, the permeability of the randomly
stacked samples for the highest Vf is lower than the
other two predicted permeabilities. Further analysis
showed that the maximum ply shift models had more
consistent paths for the uid ow; the distance which
the uid had to ow in the in-plane direction before it
reached a gap (in order to ow in the through-thickness
direction) was constant, whereas this varied in the case
of preforms with random ply shifts.
The experimental results are also shown on these
plots. These results generally lie within the range
predicted, which provides further condence in the permeability prediction tool. In the cases where the experimental values lie outside the predicted range, they are
still reasonably close despite the expected variability of
such results, previously discussed in Vernet et al.8

Journal of Reinforced Plastics and Composites 35(24)


of higher permeability correspond to locations where
the gaps between the reinforcement tows were larger.
The vertical gaps signicantly inuence the permeability in the y direction and the horizontal gaps the
permeability in the x direction. Additionally, it was
shown that the permeability is greatly aected by
local textile deformations. The approach of using average unit cells to predict the permeability of textiles was
validated: the spread of permeability results obtained
was compared with permeability results generated using
a single unit cell created from the average geometric
parameters of the textile.
The tool was also used to predict the in-plane and
through-thickness permeabilities of multi-layer preforms with increasing number of layers, with various
applied ply shifts and compacted to a range of Vfs.
These predictions mostly followed the expected trends
and were also in close agreement with the results
obtained experimentally. The tool enabled the study
of the inuence of ply shift on the permeability properties, which would otherwise be very dicult to carry
out experimentally. Most interestingly, models with a
randomly generated ply-shift led to the lowest permeability predictions for some specic cases such has high
Vf through thickness ow. This is attributed to them
having the most convoluted ow paths.
Declaration of Conflicting Interests
The author(s) declared no potential conicts of interest with
respect to the research, authorship, and/or publication of this
article.

Funding

Conclusions
A comprehensive permeability prediction tool has been
developed using textile modelling techniques. This tool
has been used to predict the permeability of a woven
reinforcing textile, capturing the inuence of geometric
variations on its permeability. The presented study
focussed on lower to medium bre volume fractions
in order to develop the principal methodologies.
Future work will focus on the predictions of models
which incorporate permeable tows and dual-scale
ows, which are expected to be more accurate at
higher volume fractions.
The inuence of geometric variations on the textiles
permeability was studied in depth. It was found that, as
expected, the parameter that has the biggest inuence
on the permeability is the unit cell tow gap width.
Complete textile permeability maps were created by
separating the textile into many individual unit cells,
predicting the in-plane permeability tensor at every
location. By superimposing the textile structure onto
the permeability maps it was shown that the locations

The author(s) received no nancial support for the research,


authorship, and/or publication of this article.

References
1. Astrom BT. Manufacturing of polymer composites. Boca
Raton, FL: CRC Press, 1997.
2. Trochu F, Ruiz E, Achim V, et al. Advanced numerical simulation of liquid composite molding for process analysis and
optimization. Compos A Appl Sci Manuf 2006; 37: 890902.
3. Trochu F, Gauvin R and Gao DM. Numerical analysis of
the resin transfer molding process by the finite element
method. Adv Polym Technol 1993; 12: 329342.
4. Darcy H. Les Fontaines Publiques de la Ville de Dijon.
Paris: Dalmont, 1856.
5. Rebenfeld L. Permeability characteristics of multilayer
fiber reinforcements. Part I: experimental observations.
Polym Compos 1991; 12: 179185.
6. Binetruy C, Hilaire B and Pabiot J. The interactions
between flows occurring inside and outside fabric tows
during RTM. Compos Sci Technol 1997; 57: 587596.
7. Arbter R, Beraud JM, Binetruy C, et al. Experimental
determination of the permeability of textiles: a benchmark
exercise. Compos A Appl Sci Manuf 2011; 42: 11571168.

Swery et al.
8. Vernet N, Ruiz E, Advani SG, et al. Experimental determination of the permeability of engineering textiles:
Benchmark II. Compos A Appl Sci Manuf 2014; 61:
172184.
9. Wong CC. Modelling the effects of textile preform architecture on permeability. Doctor of Philosophy Thesis, The
University of Nottingham, Nottingham, 2006.
10. Verleye B. Computation of the permeability of multi-scale
porous media with application to technical textiles. PhD
Thesis, Katholieke Universiteit Leuven, Leuven, 2008.
11. Lin H, Brown LP and Long AC. Modelling and simulating textile structures using TexGen. Adv Mater Res 2011;
331: 4447.
12. Zeng X, Brown LP, Endruweit A, et al. Geometrical
modelling of 3D woven reinforcements for polymer composites: prediction of fabric permeability and composite
mechanical properties. Compos A Appl Sci Manuf 2014;
56: 150160.
13. Zeng X, Endruweit A, Brown LP, et al. Numerical
prediction of in-plane permeability for multilayer woven
fabrics with manufacture-induced deformation. Compos
A Appl Sci Manuf 2015; 77: 266274.
14. Swery EE, Meier R, Lomov S, et al. Predicting permeability based on flow simulations and textile modelling
techniques; comparison with experimental values and
verification of FlowTex solver using Ansys CFX.
J Compos Mater 2015; 50: 601615.
15. Verleye B, Croce R, Griebel M, et al. Permeability of
textile reinforcements: simulation, influence of shear
and validation. Compos Sci Technol 2008; 68: 28042810.
16. Wong CC, Long AC, Sherburn M, et al. Comparisons of
novel and efficient approaches for permeability prediction based on the fabric architecture. Compos A Appl
Sci Manuf 2006; 37: 847857.
17. Ngo ND and Tamma KK. Complex three-dimensional
microstructural permeability prediction of porous fibrous
media with and without compaction. Int J Numer Method
Eng 2004; 60: 17411757.
18. Loix F, Badel P, Orgeas L, et al. Woven fabric permeability: from textile deformation to fluid flow mesoscale
simulations. Compos Sci Technol 2008; 68: 16241630.
19. Griebel M and Klitz M. Homogenization and numerical
simulation of flow in geometries with textile microstructures. Multiscale Model Simulat 2010; 8: 14391460.
20. Swery EE, Allen T and Kelly PA. Automated tool to determine geometric measurements of woven textiles using digital image analysis techniques. Text Res J. Epub ahead of
print 15 July 2015. DOI: 10.1177/0040517515595031.
21. Bickerton S and Gan JM. Accounting for reinforcement variability in liquid moulding, utilising optically measured areal
weight maps. In: SAMPE, Seattle, WA, 1720 May 2010.
22. Desplentere F, Lomov SV, Woerdeman DL, et al. MicroCT characterization of variability in 3D textile architecture. Compos Sci Technol 2005; 65: 19201930.
23. Abdiwi F, Harrison P, Koyama I, et al. Characterising
and modelling variability of tow orientation in engineering fabrics and textile composites. Compos Sci Technol
2012; 72: 10341041.
24. Endruweit A, Zeng X and Long A. Effect of specimen
history on structure and in-plane permeability of woven

1813

25.

26.

27.

28.

29.

30.

31.

32.

33.

34.

35.

36.

37.

38.

39.

40.

fabrics. J Compos Mater. Epub ahead of print 22 May


2014. DOI: 10.1177/0021998314536070.
Rieber G, Jiang J, Deter C, et al. Influence of textile
parameters on the in-plane Permeability. Compos A
Appl Sci Manuf 2013; 52: 8998.
Swery EE, Allen T and Kelly PA. Predicting compaction
induced deformations of meso-scale textile models efficiently. J Compos Mater. In press.
Drago A and Pindera M-J. Micro-macromechanical analysis of heterogeneous materials: macroscopically homogeneous vs periodic microstructures. Compos Sci Technol
2007; 67: 12431263.
Ivanov I and Tabiei A. Three-dimensional computational
micro-mechanical model for woven fabric composites.
Compos Struct 2001; 54: 489496.
Takano N, Zako M, Okazaki T, et al. Microstructurebased evaluation of the influence of woven architecture
on permeability by asymptotic homogenization theory.
Compos Sci Technol 2002; 62: 13471356.
Staffan Lundstrom T. Numerical study of the local permeability of noncrimp fabrics. J Compos Mater 2005; 39:
929947.
Belov EB, Lomov SV, Verpoest I, et al. Modelling of
permeability of textile reinforcements: lattice Boltzmann
method. Compos Sci Technol 2004; 64: 10691080.
Potluri P and Sagar TV. Compaction modelling of textile
preforms for composite structures. Compos Struct 2008;
86: 177185.
Bayraktar H, Ehrlich D, Scarlat G, et al. Prediction of
yarn paths in 3D woven composites using finite element
simulations. In: Lomov S (ed.) TexComp-11: international
conference on textile composites, Leuven, 1517
September 2013.
Lin H, Sherburn M, Crookston J, et al. Finite element
modelling of fabric compression. Model Simulat Mater
Sci Eng 2008; 16: 035010.
Tavana R, Shaikhzadeh Najar S, Tahaye Abadi M, et al.
Meso/macro-scale finite element model for forming process of woven fabric reinforcements. J Compos Mater
2013; 47: 20752085.
Govignon Q, Bickerton S and Kelly PA. Simulation of
the reinforcement compaction and resin flow during the
complete resin infusion process. Compos A Appl Sci
Manuf 2010; 41: 4557.
Swery EE, Allen T, Cardona C, et al. Efficient experimental characterisation of the permeability of fibrous textiles.
J Compos Mater. Epub ahead of print 22 February 2016.
DOI: 10.1177/0021998316630801.
Liotier P-J, Govignon Q, Swery E, et al. Characterisation
of woven flax fibres reinforcements: effect of the shear on
the in-plane permeability. J Compos Mater. Epub ahead of
print 14 January 2015. DOI: 10.1177/0021998314565411.
Comas-Cardona S, Cosson B, Bickerton S, et al. An optically-based inverse method to measure in-plane permeability fields of fibrous reinforcements. Compos A Appl
Sci Manuf 2014; 57: 4148.
Zeng X, Long A and Endruweit A. Numerical prediction
of permeability of textiles for the international benchmark
exercise. In: FPCM 11: the 11th international conference on
flow processes in composite materials, Auckland, 2012.

Vous aimerez peut-être aussi