Download as pdf or txt
Download as pdf or txt
You are on page 1of 125

ETH Library

Numerical Analysis and Modelling


of the Velocity Distribution of
Saturated Resin Flow During LCM

Master Thesis

Author(s):
Krüsi, Stephan

Publication date:
2015

Permanent link:
https://doi.org/10.3929/ethz-a-010564904

Rights / license:
In Copyright - Non-Commercial Use Permitted

This page was generated automatically upon download from the ETH Zurich Research Collection.
For more information, please consult the Terms of use.
Master’s Thesis

Numerical Analysis and Modelling of


the Velocity Distribution of Saturated
Resin Flow During LCM

Author: Supervisor:
Stephan Krüsi Jesús C. Maldonado

Prof. Dr. Paolo Ermanni


Laboratory of Composite Materials and Adaptive Structures
Department of Mechanical and Process Engineering

Reference Number: 15-020 April 2015


ETH Zürich
IDMF - Laboratory of Composite Materials and Adaptive Structures
LEE O 203
Leonhardstrasse 21
8092 Zürich

Telephone: +41 (0)44 633 63 02

www.structures.ethz.ch
Abstract

Fiber reinforced composite materials offer a high degree of optimization for advanced
structures. A solid understanding of the physics involved in the processing of polymer
composites is equally as important as the material properties of the fibers and matrix
themselves. Dual-scale fabrics are commonly employed in liquid composite molding
processes. These type of porous fabrics give rise to dual-scale flows, which deviate
strongly from the homogeneous flow field commonly described using Darcy’s Law.

This thesis develops a numerical framework for analyzing saturated, dual-scale flows in
composite textiles. Using the software TexGen, 3D models of the textiles are generated
to produce repeating unit cells. A fluid dynamics simulation of the flow through the
dual-scale textile is implemented in COMSOL Multiphysics. The Navier-Stokes and
Stokes-Brinkman equations are applied in inter- and intra-tow flow regimes respectively.
The numerical analysis of the velocity field shows that the inter-tow flow has a non-
uniform distribution. A study of the volumetric flow rates demonstrates that a majority
of the mass transport occurs in a highly localized region of the inter-tow flow channels.
The flow rate analysis also shows potential for developing an improved dual-scale flow
model, which includes the variations of the the inter-tow flow.
Zusammenfassung

Faserverstärkte Kunststoffe ermöglichen einen hohen Optimierungsgrad komplexer Struk-


turen. Ebenso wichtig wie die Materialeigenschaften des Verbundwerkstoffes ist ein
fundiertes Verständnis der Physik und der Herstellungsprozesse. Sogenannte Zwei-
Skalen-Gewebe werden oft in Liquid-Composite-Molding-Prozessen eingesetzt. Derar-
tige Textilien haben zwei unterschiedliche Grössenordnungen von Porosität (Leerräume)
die ein inhomogenes Fliessfeld verursachen, welches nicht mit dem Darcy-Gesetz be-
schrieben werden kann.

Im Rahmen dieser Masterarbeit wurde ein numerisches System entwickelt, welches die
zweiskaligen Flüsse in Faserverbundwerkstoffen analysiert. Die dreidimensionalen Ge-
ometrien der Gewebe wurden mit der Software TexGen nachgebildet, um periodische
Einheitszellen zu generieren. Die Fluiddynamik-Simulation wurde in COMSOL Multi-
physics implementiert und berücksichtigt die Zwei-Skalen-Effekte. Navier-Stokes- und
Stokes-Brinkman-Gleichungen repräsentieren die Flüsse zwischen und innerhalb den
Faserbündeln. Die numerische Analyse des Geschwindigkeitsfeldes zeigt grosse Vari-
ationen im Fluss zwischen den Bündeln. Die Analyse des Volumenstroms weist da-
rauf hin, dass ein massiver Anteil des Massentransports innerhalb lokalisierter Bereiche
in den Zwischenfaserbündel-Kanälen stattfindet. Anhand dieser Volumenstromanalyse
wird ein verbessertes mathematisches Modell vorgeschlagen, welches den Einfluss der
Zwei-Skalen-Effekte abbildet.
Acknowledgements

The work in this thesis could have by no means been done alone. My first and foremost
thanks goes to Jesús Maldonado. Thank you for being a great thesis advisor, for your
insightful discussions (on and off topic) and sound advice when I was stuck. I would
also like to thank Prof. Ermanni, Dr. Kress, and the PhD students at the CMAS Lab
for your constant willingness to help us students. You have all made working at our
department a great experience.

A special thanks also goes to my friends who not only offered creative ideas for my
research, but transformed the challenges of studying at ETH into a very enjoyable
experience. Finally, my thanks and love goes to my family who, although spread out
across the globe, are always there supporting and encouraging me.
Contents

Abstract ii

Contents v

List of Figures viii

List of Tables x

Nomenclature xi

Abbreviations xii

1 Introduction 1
1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Composite Materials and Processing . . . . . . . . . . . . . . . . . . . . 3
1.2.1 Historical Background . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2.2 Thermoset Composite Processing . . . . . . . . . . . . . . . . . . 5
1.2.3 Computational Modelling . . . . . . . . . . . . . . . . . . . . . . 7
1.3 The State of the Art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.1 Dual-Scale Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.3.2 Numerical Modelling of Resin Flow . . . . . . . . . . . . . . . . . 9
1.3.3 Unit Cell Modelling . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4.1 Potential Applications . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4.2 Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2 Dual-Scale Textiles 14
2.1 Material Characterization . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.1 Geometrical Properties . . . . . . . . . . . . . . . . . . . . . . . 15
2.1.2 Dual-Scale Fabrics . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.1.3 Ethylcinnamate . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.2 3D Modelling of Textiles . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2.1 TexGen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.2.2 Generating a CFD Unit Cell . . . . . . . . . . . . . . . . . . . . 24
2.2.3 Resulting Textiles . . . . . . . . . . . . . . . . . . . . . . . . . . 28

3 Dual-Scale Flow 29
3.1 Porous and Free Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

v
Contents vi

3.1.1 Darcy’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30


3.1.2 The Navier-Stokes Equations . . . . . . . . . . . . . . . . . . . . 31
3.2 Micro-scale Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2.1 Intra-tow Permeability Models . . . . . . . . . . . . . . . . . . . 33
3.2.2 Intra-tow Permeability Simulation . . . . . . . . . . . . . . . . . 36
3.2.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
3.3 Meso-scale Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.1 Intra-tow flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.3.2 Inter-tow flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

4 Solving the Dual-Scale Flow 43


4.1 COMSOL Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.1.1 Geometry Import and Input Parameters . . . . . . . . . . . . . . 44
4.1.2 Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
4.1.3 Flow Equations and Element Order . . . . . . . . . . . . . . . . 46
4.1.4 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . 47
4.1.5 Solver Settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.2 Meshing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.2.1 Convergence Study . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.3 EULER High Performance Cluster . . . . . . . . . . . . . . . . . . . . . 53
4.3.1 Running COMSOL Simulations on EULER . . . . . . . . . . . . 53
4.4 Data Extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

5 Results and Discussion 57


5.1 Velocity Distribution in a Plain Weave . . . . . . . . . . . . . . . . . . . 57
5.1.1 Histograms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
5.1.2 Effect of Fiber Volume Content . . . . . . . . . . . . . . . . . . . 60
5.1.3 Volumetric Flow Rates . . . . . . . . . . . . . . . . . . . . . . . . 63
5.2 Comparison to a Biaxial Fabric . . . . . . . . . . . . . . . . . . . . . . . 65
5.2.1 CFD Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.2.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.3 Comparison to Experiments . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.3.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.3.2 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
5.4 Comparison with the State of the Art . . . . . . . . . . . . . . . . . . . 72
5.4.1 Previous Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
5.4.2 Dual-Scale Flow Simulation . . . . . . . . . . . . . . . . . . . . . 72
5.5 Equivalent Flow Channel Model . . . . . . . . . . . . . . . . . . . . . . 76

6 Conclusions and Outlook 78


6.1 Further Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
6.2 Potential Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Bibliography 81
Contents vii

Appendix A Official Documents 87

Appendix B Unit Cell Generator Code 93

Appendix C Calculation of Pressure Gradient 105

Appendix D Datasheets 107


List of Figures

1.1 Overview of thesis work . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Percentage of materials used in a Boeing 787 body . . . . . . . . . . . . 4
1.3 Boeing’s substantial increase in CFRP usage . . . . . . . . . . . . . . . 4
1.4 A schematic of an RTM and Compression-RTM process . . . . . . . . . 6
1.5 The micro, meso, and macro scales within a fiber preform . . . . . . . . 8
1.6 Flow simulations on the meso and macro scales . . . . . . . . . . . . . . 10
1.7 Evident dual-scale flow in a two-color resin injection . . . . . . . . . . . 13

2.1 Different types of dual-scale textiles . . . . . . . . . . . . . . . . . . . . 14


2.2 Microscopy image of a woven aramid composite laminate . . . . . . . . 15
2.3 Geometrical parameters needed to describe the yarns in a fabric . . . . 16
2.4 Various cross-sections generated using the power ellipse function . . . . 16
2.5 Reduction in fabric height due to compaction pressure . . . . . . . . . . 17
2.6 Close-up images of two types of glass fabrics . . . . . . . . . . . . . . . . 19
2.7 Close-up image of a UD carbon fabric . . . . . . . . . . . . . . . . . . . 20
2.8 Dynamic viscosity measurements of the Ethylcinnamate . . . . . . . . . 22
2.9 A 3D orthogonal fabric generated with TexGen . . . . . . . . . . . . . . 23
2.10 Flow chart of the unit cell generator implemented in Python/TexGen . 25
2.11 Stages for generating a periodic unit cell . . . . . . . . . . . . . . . . . . 27

3.1 Darcy’s original sand column experiment . . . . . . . . . . . . . . . . . . 30


3.2 Micro-scale model: an array of hexagonally packed cylinders . . . . . . . 33
3.3 Axial and tranverse permeability simulations (Vf = 0.60) . . . . . . . . . 37
3.4 Permeability Kk for increasing Vf - Longitudinal Flow . . . . . . . . . . 39
3.5 Permeability K⊥ for increasing Vf - Transverse Flow . . . . . . . . . . . 39
3.6 Intra- and inter-tow flow regions in a meso-scale unit cell . . . . . . . . 40

4.1 Flow chart for the numerical analysis of dual-scale velocity distributions 43
4.2 Automated selections for various model parameters . . . . . . . . . . . . 45
4.3 Flow-driving pressure drop (Pa) over unit cell . . . . . . . . . . . . . . . 47
4.4 Boundary conditions applied to top & bottom and left & right walls . . 48
4.5 The mesh of the dual-scale unit cell . . . . . . . . . . . . . . . . . . . . 49
4.6 Comparison of three meshing strategies . . . . . . . . . . . . . . . . . . 50
4.7 Convergence study for plain weave fabric (Vf = 0.33, ∆p = 200 Pa) . . . 51
4.8 A zoomed-in view of the regular grid . . . . . . . . . . . . . . . . . . . . 55

5.1 Resulting flow field of the dual-scale simulation . . . . . . . . . . . . . . 57


5.2 Velocity distribution for plain weave fabric (Vf = 0.40) . . . . . . . . . . 58
5.3 Distribution of the cross-sectional area for increasing velocity regimes . 60

viii
List of Figures ix

5.4 Velocity distribution for intra-tow domain . . . . . . . . . . . . . . . . . 61


5.5 Velocity distribution for inter-tow domain . . . . . . . . . . . . . . . . . 62
5.6 Comparison of area occupied vs. flow rates for the plain weave fabric . . 64
5.7 Flow rate regimes compared to area occupied for multiple Vf . . . . . . 64
5.8 A biaxial fabric generated with TexGen . . . . . . . . . . . . . . . . . . 65
5.9 Mesh and resulting flow field of biaxial fabric (Vf = 0.44) . . . . . . . . 66
5.10 Biaxial fabric: velocity distributions for inter-tow domain . . . . . . . . 67
5.11 Biaxial fabric: flow rate vs. area comparison for multiple Vf . . . . . . . 67
5.12 RTM tool developed for dual-scale flow tracking . . . . . . . . . . . . . . 68
5.13 Comparison between absolute velocities . . . . . . . . . . . . . . . . . . 70
5.14 Comparison between normalized velocity distributions . . . . . . . . . . 70
5.15 Visualization of the area selection around flow channels . . . . . . . . . 73
5.16 Comparison between dual-scale flow rate models . . . . . . . . . . . . . 73
5.17 Comparison between dual-scale flow rate models and biaxial simulation 74
5.18 Comparison between the plain weave and biaxial simulations . . . . . . 74

C.1 Schematic of RTM injection experiment . . . . . . . . . . . . . . . . . . 106


C.2 Calculated pressure drops over the experiment components . . . . . . . 106
List of Tables

2.1 Plain weave glass fabric properties . . . . . . . . . . . . . . . . . . . . . 19


2.2 Twill 2/2 glass fabric properties . . . . . . . . . . . . . . . . . . . . . . . 19
2.3 Carbon fiber UD fabric properties . . . . . . . . . . . . . . . . . . . . . 21
2.4 Ethlycinnamate (CAS-Nr 103-36-6) material properties . . . . . . . . . . 21
2.5 Generated plain weave and biaxial unit cells . . . . . . . . . . . . . . . . 28

3.1 Constants used for the permeability formulas with hexagonal packing . . 35
3.2 Micro-scale simulation parameters . . . . . . . . . . . . . . . . . . . . . 37

4.1 Meso-scale simulation parameters . . . . . . . . . . . . . . . . . . . . . . 45


4.2 Computational workload on BRUTUS and EULER clusters . . . . . . . 54
4.3 Dimensions for the plain weave and biaxial regular grids . . . . . . . . . 56

5.1 Materials used in the particle tracking experiment . . . . . . . . . . . . 69

x
Nomenclature

ρ fluid density kg/m3


µ dynamic viscosity Pa s
L characteristic length m
Re Reynolds number -
Q volumetric flow rate m3 /s
Qintra intra-tow volumetric flow rate m3 /s
Qinter inter-tow volumetric flow rate m3 /s
A area m2
An area of single grid point m2
Av area of binned velocity range m2
P pressure N/m2
p pressure in FEM N/m2
K permeability tensor m2
K permeability m2
Kk permeability parallel to fibers m2
K⊥ permeability perpendicular to fibers m2
v fluid velocity m/s
ṽav average porous flow velocity m/s
ṽP T V average empty cavity velocity m/s
v velocity vector m/s
f external forces vector N
u velocity vector in FEM m/s
I identity matrix -
p porosity -
Vf Fiber volume content (fabric) -
Vf -tows Fiber volume content (tows) -
TG glass transition temperature ◦C

xi
Abbreviations

BC Boundary Condition
CAD Computer Aided Design
CFD Computational Fluid Dynamics
CFRP Carbon Fiber Reinforced Polymer
CPU Central Processing Unit
ETHZ Eidgenossische Technische Hochschule Zürich
FVC Fiber Volume Content
FTP File Transfer Protocol
GUI Graphical User Interface
HPC High Performance Cluster
LCM Liquid Composite Molding
LSF Load Sharing Facility
NF Nesting Factor
NCF Non-Crimped Fabric
OOA Out-of-Autoclave
PTV Particle Tracking Velocimetry
RAM Random Access Memory
REV Representative Volume Element
RUC Repeating Unit Cell
RTM Resin Transfer Molding
UD Uni-Directional (fabric)
VARI Vacuum Assisted Resin Infusion

xii
Chapter 1

Introduction

1.1 Overview

This thesis investigates the nature of polymeric resin flows through the reinforcing fabrics
that are commonly used in composite materials. The work is primarily numerical,
meaning that fluid dynamics simulations are carried out on 3D models of the textiles.
The resulting velocity fields are analyzed in order to study the effects of the textile
architecture on the local resin flows.

Within the scope of this thesis, a framework for analyzing dual-scale flows has been
implemented. The material properties of the fabric are incorporated into the generation
of an appropriate textile model. Then the dual-scale flow is simulated and an insightful
analysis of the resulting data is performed. With this output, the influences of the fabric
architecture on the velocity distributions and resin transport can be studied.

Dual-scale flow analysis framework

Textile
Generator Results:

Material CFD • Fabric architectures


Characterization Simulation • Velocity distributions
• Mass transport
Data Analysis

Figure 1.1: Overview of thesis work

1
Chapter 1: Overview 2

The structure of the thesis is outlined as follows:

Chapter 1 gives a general introduction to the history of composite materials, their


applications, and various fabrication techniques. The current state of scientific research
relevant to this thesis is presented. Then the potential applications, motivation, and
goals of this thesis are outlined.

Chapter 2 introduces the notion of dual-scale textiles. A material characterization of


textiles used in this thesis is given. The method with which 3D models of the textiles
are generated is explained in detail.

Chapter 3 describes the flow of resin on the micro- and meso-scale. For the micro-scale
flow, various models for determining the permeability within a fiber are tow presented
and compared to simulated results. A coupling strategy for the micro-mesco-scale sim-
ulation is introduced, and the flow equations for each regime are explained.

Chapter 4 outlines the flow simulation. Each part of the fluid dynamics model and how
it is built up is described. The boundary conditions, flow equations, meshing strategy,
and solver settings are explained. The chapter ends by describing the computational
environment and data extraction method.

Chapter 5 presents the analysis performed on the dual-scale flow data. The velocity
distribution and mass transport of the intra- and inter-tow regions are shown and ex-
plained. A comparison of two types of fabrics is made, as well as a comparison to results
from a dual-scale flow experiment. Finally, the results of this thesis are compared to
state of the art models.

Chapter 6 ends with an interpretation of the results presented in this thesis. Im-
provements to the geometrical and fluid dynamics models are suggested, and potential
applications of the flow rate model are explored.
Chapter 1: Composite Materials and Processing 3

1.2 Composite Materials and Processing

1.2.1 Historical Background

Mankind’s use of composite materials is evident throughout our recorded history. The
realization that materials could be strengthened by adding a second reinforcing material
was made very early on. Wooden tools from the Swiss lake-dwellers, which purposely
made use of the natural fiber reinforcement in the wood, have been found dating back to
5000 BC [1]. The Israelites produced strengthened mud bricks by adding straw to them.
And the Egyptians, by rearranging and glueing together wood with different grain-
directions, improved its strength and resistance to swelling [2]. Even though the addition
of reinforcing materials adds a level of complexity compared to isotropic (directionally
uniform) materials, the gained benefits have established composite materials as a worth
while endeavour.

Carbon fiber reinforced polymers (CFRPs) are the pinnacle of composite material tech-
nology. Carbon fibers were first used by Thomas Edison in 1879 as the conducting
filament for his early light bulbs, but it was not until the post World War II era that
industrial use of carbon fibers began to take off [3, 4]. The lighter-than-steel, yet strong
and stiff composite materials opened up new possibilities in designing light-weight struc-
tures. The use of CFRPs in advanced structures has steadily increased since their broad
introduction in the 1940’s [5, 6].

In the 1960’s, new production techniques led to the use of rayon and PAN (polyacryloni-
trile) fibers which boasted very high strength and high modulus (stiffness) properties. It
is no coincidence that during the 1960’s the development and subsequent use of carbon
fiber materials was primarily backed by governments around the world. Military aircraft
and spacecraft needed ever increasingly sophisticated materials, whose demand CFRPs
were able to meet. In addition to being light-weight and strong, carbon fibers showed
higher thermal stability than glass fibers, which made them ideally suited for rocket
nozzle exit cones and re-entry heat shields [3]. By the 1970’s many US fighter jets were
using CFRPs for increasingly significant parts of the aircraft structure, which paved the
way for commercial aircraft adoption beginning during the 1980’s [1, 5].

Rise in Industrial Applications

Over the past four decades, composites structures have revolutionized structural designs
in commercial aviation. At first only minor elements, such as the door springs of the
Boeing 767, were made out of composites. Then primary structures like the tail rudder
Chapter 1: Composite Materials and Processing 4

Figure 1.2: Percentage of materials used in a Boeing 787 body [9]

Figure 1.3: Boeing’s substantial increase in CFRP usage [7]

and elevators began to be replaced with CFRPs [7]. Both Boeing and Airbus made
these early adoptions, and with their most recent flagship airliners, the 787 Dreamliner
and Airbus A380, brought the use of composite materials in aviation to a whole new
dimension. Figure 1.3 shows the ever increasing trend in composite usage within the
Boeing fleet, with the 787 Dreamliner using 50% CFRPs by weight [8, 9].

The primary motivation to increase the use of CFRPs in aircraft structures has been
cost savings. Weight reductions of 20-50% significantly reduce the fuel consumption,
saving costs over the entire service lifetime. Composite materials also offer other specific
benefits that appeal to composite usage in aerospace applications. During the plane’s
construction, assembly time as well as weight can be reduced using integral construction
techniques offered by composite structures [10]. Also, CFRPs show a much higher
resistance to fatigue and corrosion than aluminium [11], the traditional material used
for aircraft structures. Because of this, aircraft using composite materials promise to
require less planned and non-routine maintenance, again leading to significant long-term
cost reductions [8].
Chapter 1: Composite Materials and Processing 5

The space industries also profited from the superior properties of glass, carbon, and
ceramic composite materials. The development of pitch-based, ultrahigh-modulus car-
bon fibers in the late 1970’s became very useful for space satellites, which demanded
very rigid structures even when facing temperature differences of over 200°C [12]. To-
day, NASA and ESA make extensive use of CFRPs for their launch vehicles, satellite
structures, and space telescopes [13].

1.2.2 Thermoset Composite Processing

There are many ways to manufacture fiber reinforced polymer composites. There is
hand lamination, where resin is rolled onto the dry fibers by hand, prepreg processes,
where the fibers are pre-impregnated with resin and then cured as one piece, and infusion
methods, where a liquid resin is pushed or drawn through a dry fabric. Each of these
processing variants has advantages and disadvantages which make them suitable for
different applications [1, 5].

Hand lamination offers flexibility and low costs as it does not require expensive tooling,
but suffers from poor part consistency, high void content, and a low upper limit to the
achievable fiber volume content (Vf ≈ 50%). Prepreg technology, on the other hand, is
the premium technique in composite processing. Repeatably able to produce the highest
quality parts (Vf > 60%, void content < 2%), prepregs have for a long time been the
preferred process for the aerospace industry [11, 14]. However, prepreg technology is
also premium in its costs; besides high material costs, the partially cured prepregs must
be kept in cooled storage facilities until they are used, and the initial investments for
the required tooling (autoclaves) are immense. A NASA cost study [15] found that it
would cost $40 million to build and $60 million to install an autoclave large enough to
produce launch vehicle parts.

Due to the high costs of autoclave processes, much interest and research has been in-
vested in out-of-autoclave (OOA) prepreg technology. OOA prepregs offer the same
reproducibility and high material property attributes as traditional in-autoclave pro-
cesses, but avoid the high costs and long cycle times related to the autoclave itself
[14, 15]. By fine tuning their material and processing parameters, OOA prepregs can
also mitigate voids enough to produce aerospace-grade parts [11].

Liquid Composite Molding

As their name implies, Liquid Composite Molding (LCM) processes involve the injection
of a liquid resin into a dry fiber bed. The composite part, typically made of a polymeric
Chapter 1: Composite Materials and Processing 6

Figure 1.4: A schematic of an RTM and Compression-RTM process [16]

resin and carbon, glass, or aramid fibers, is then formed and cured within a mold. There
can be many variants to this general process, which has given rise to a whole range of
processes included within the LCM family: Resin Transfer Molding (RTM), Vacuum
Assisted Resin Transfer Molding (VARTM), Vacuum Assisted Resin Infusion (VARI),
and Resin Film Infusion (RFI), to name a few [1, 16]. Each of these processes has certain
advantages and are all found producing specialized parts for the aerospace industry [6].

A typical RTM process involves the injection of resin into a closed mold. The resin can be
heated during infusion in order to reduce its viscosity and speed up the injection, or it can
be cured after being fully injected. Figure 1.4 shows another RTM variant: Compression-
RTM (CRTM). Here the mold is held slightly open during injection (increasing the
permeability of the fiber bed) and then compresses once the fibers are fully wetted with
resin. This variant is aimed at reducing the time needed for the resin injection, while
still ensuring a good quality part [16].

The closed mold solution of RTM processes allows for the resin to be injected at high
pressures (up to 100 bar cavity pressure), which reduces the cycle time for each injection
[17]. This advantage makes RTM the process of choice for medium-size series production
(5-50 thousand parts per year) [1]. RTM processes can even produce sizable parts in
just minutes [18], compared to the many hours required by an autoclave cycle. Recently,
state of the art LCM techniques have been able to produce aerospace-grade parts with
Vf ≈ 60% and void content of < 2%, matching prepregs [6].

In contrast to RTM, where resin is pushed through the part by an applied pressure,
Vacuum Assisted Resin Infusion (VARI) pulls the resin through the fiber bed by drawing
vacuum at the resin outlet. Not only does this eliminate the need for high-pressure
pumps, but greatly reduces the cost of the mold as it does not need to be as stiff
(to resist deformation under the high resin pressures in RTM) [11, 19]. As a further
Chapter 1: Composite Materials and Processing 7

cost reduction, VARI molds are one-sided, as the atmospheric pressure around the part
presses down on the saturated fibers, giving the part its desired shape. Low-cost tooling
make VARI the ideal processes when making large parts, where the cost and physical
requirements of an RTM mold would be too high. Large structures for ships (hulls,
decks) [20] or wind turbine blades are commonly made with VARI processes [19].

1.2.3 Computational Modelling

The advent of high-performance computers has revolutionized many engineering disci-


plines such as structural analysis, fluid dynamics, and product design. The way com-
posite structures are designed and manufactured has also seen significant improvement
in the last decades due to improved modelling capabilities [21]. By implementing nu-
merical methods, the composite engineer can produce increasingly complex parts while
reducing the design-to-market time.

Simulation techniques have not only improved the structural analysis of composites,
but also their processing methods. Newer models take into account void formation and
resin flows through complex reinforcing fabrics [22]. The final quality of a composite
part also strongly depends on the resin properly filling and curing within the mold
[23]. Traditionally, designing the mold and finding the optimal resin injection strategy
was largely based on a trial-and-error approach and on the experience of the engineer.
This time-consuming approach can be reduced by simulating the resin injection and
numerically optimizing the mold design before producing a costly prototype [24, 25].

A better understanding of reality, paired with a more accurate numerical model, allows
for a highly optimized structure (in both cost and functionality) to be built.
Chapter 1: The State of the Art 8

1.3 The State of the Art

This section describes some of the foundational ideas and current research most relevant
to the topic presented in this thesis: modelling resin flow through reinforcing fabrics.

Section 1.2 showed that LCM processes make up a significant part of the composites
industry. The ability predict and simulate key process parameters, such as filling time,
injection temperature and port locations, not only reduces production costs but signif-
icantly improves the quality of parts [16, 23, 26]. Both physically sound models and
robust simulation techniques are required if such predictions are to be applied in indus-
try.

Current resin injection software packages (PAM-RTM, RTM-Worx, LIMS) simulate the
mold filling process using Darcy’s Law [25, 27]. The key parameter for these simulations
is the textile permeability, which is why much of the current LCM modelling research
focuses on determining permeabilities for varying preform parameters (fiber volume
content, fiber architecture) [23, 25, 28].

1.3.1 Dual-Scale Flow

Verleye [29] and Tan and Pillai [24] describe porous flow on three different levels: the
micro-scale (within the fiber tow [µm]), the meso-scale (a single unit cell of a fiber
network [mm]), and the macro-scale (the entire composite part [m]). This distinction
between the micro- and meso-scale flow is important, as the resistance to flow is orders
of magnitude higher within the tows. In [30], Pillai states that the “resin inside the tows
is practically stationary with respect to the resin traveling through the gaps around the
tows”.

Pillai’s 2002 work [26] laid the foundational theory for modelling dual-scale flow in
porous textiles. The theory describes dual-scale, unsaturated/two-phase flow, which is
flow on the transitional region between air and liquid phases. Previously only the flow

Figure 1.5: The micro, meso, and macro scales within a fiber preform [31]
Chapter 1: The State of the Art 9

front was considered two-phase, but Parnas and Phelan [32] had shown that because of
the dual-scale nature of fabrics, an unsaturated region can occur behind the flow front.
By adding a negative sink term to the continuity equation, Pillai could account for the
resin loss into the not yet fully saturated tows within the fabric.

Many dual-scale effects that arise in unsaturated flow have been studied (creation of
voids [22], droop in inlet pressure [33]). It would be a logical step to investigate the dual-
scale velocity distribution in a unit cell with unsaturated flow. However, the analysis
performed in this thesis does not consider two-phase, unsaturated flow. It models only
fully saturated flow. Nevertheless, a fully-saturated investigation does show clear dual-
scale effects and describes how the resin flows very close to the unsaturated flow front
without adding the complexity of modelling transient, unsaturated flow.

A key element to current dual-scale models [26, 29] is the volume averaging method.
This method homogenizes the physical flow properties (pressure or velocity for example)
in the considered volume:

Z
1
hvg i = vg dV (1.1)
V Vg

where V is the volume to be averaged over, and Vg and vg are the volume and velocity in
the free flow gaps. With this volume averaging method, effective velocities are calculated
for each regime of the dual-scale medium [24]. These volume averaged properties become
parameters for a macro-scale Darcy’s Law, which can be used for LCM filling simulations
of real parts. There is a trade-off between having a numerically feasible simulation and
loosing some information about the dual-scale flow within the fabric.

1.3.2 Numerical Modelling of Resin Flow

Computational Fluid Dynamics (CFD) simulations are notorious for being computa-
tionally expensive, to the extent that they often become cost or time prohibitive for a
given application. Modelling the resin injection on the scale of a whole composite part
(macro) by means of CFD simulation is not computationally feasible. To remedy this,
the practice of simulating the fluid flow over micro- and/or meso-scale repeating unit
cells (RUCs) dominates current research.

Ngo and Tamma [28] use a volume-averaged Stokes-Brinkman equation to model the
saturated flow through a meso-scale unit cell. Using the parameter α as a switch, they
apply a single flow equation in the regions within and outside of the fiber bundles. They
Chapter 1: The State of the Art 10

[mm] scale [m] scale

Figure 1.6: Flow simulations on the meso and macro [27] scales

couple the micro- and meso-scale flow equations by calculating the inter-tow permeability
with Gebart’s semi-analytical model [34].

Wong et al. [23] focus on permeability calculations. Because a large number of flow sim-
ulations are needed to account for textile variations, they developed the Grid2D solver.
This novel approach reduces the 3D flow simulation to 2D and calculates the pressure
field in the RUC with a finite difference solver. The permeability values extracted from
the Grid2D method compare well to those calculated from a full CFD simulation, show-
ing its usefulness when solving for textile permeabilities. The authors use their software
TexGen, which is able to build 3D models of a wide range of fabric preforms. The
TexGen software is explained in detail in Section 2.2.

Verleye et al. [25, 29] extend the previous work with a 3D Stokes solver, NaSt3DGP,
which solves the Stokes equations on a 3D regular grid using a finite difference discretiza-
tion. Discretizing the complex fabric geometry into small voxels (volume elements) al-
lows for a robust meshing process, regardless of fabric architecture. In regions where the
intra-tow flow is calculated, they apply the Brinkman equations. The authors propose
coupling their meso-scale solver with an existing macro-scale Darcy solver to simulate
the entire impregnation process.

Tan and Pillai [24] model unsaturated flows using a finite element/control volume
scheme. They first calculate the permeability of a unit cell, and then apply it to the
transient filling simulation of the macro-scale fabric. Wang and Grove [21] follow a
similar approach in calculating volume averaged flows on the meso-scale.

The specific implementation of the flow simulation varies based on the flow state consid-
ered (saturated vs. unsaturated) and information sought (permeability vs. flow field).
All of these methods show that simulating flows through the meso-scale structure of
fabrics can give accurate information about LCM process parameters.
Chapter 1: The State of the Art 11

1.3.3 Unit Cell Modelling

A geometrical representation of the dual-scale textile is needed to simulate the flow


of resin. Composite preforms have a repeating textile architecture, which lends itself
to being modelled with periodic unit cells. The input parameters needed to create a
representative geometry depend on the fiber type, weave patter, and yarn shape and
dimensions. The repeating unit cell (RUC) is the most elementary representation of
the fabric: one unit cell fully describes the idealized fabric. However, processing pa-
rameters such as the compaction pressure or draping of the preform cause changes and
irregularities in the fabric. In reality, deviations to the idealized RUC exist [29].

The cross-sectional shape of the fabric yarns is influenced by many factors: number
of fiber filaments in the tows, tow width, the applied compaction pressure, and fabric
architecture (weave pattern). Early efforts [35] modelled the tows as cylinders, but this
is a gross simplification.

Pillai [30] describes woven yarn cross-sections as being close to elliptical, Vandeurzen
[36] and Sherburn [37] as lenticular. Crookston [38] uses a flexible function, which allows
for changes in the aspect ratio and tow shape. The shape of the yarn cross-section can
also change during processing. Chen & Chou [39] demonstrate the height reduction of
yarns as a function of the fiber volume fraction.
Chapter 1. Motivation 12

1.4 Motivation

1.4.1 Potential Applications

What is the usefulness, or added value of a detailed model of the dual-scale flow? Un-
derstanding from and to where the resin flows during an injection allows for a better
optimization of the process parameters. Particularly when complex resin systems (multi-
phase or those including nano- or micro-particles) are used, the effects of the fabric on
the resin flow become increasingly important.

Particle-filled resins are commonly employed to increase the toughness, strength, or


electrical properties of the composite part [40]. The advantage of using a particle-filled
resin is that the improvements to the part are integrated into the existing process of the
resin injection. These processes encounter problems such as particle deposition or cake
filtering, which arise from the interactions of the resin with the dual-scale fabric [41].

Knowing when and how much of the particle-filled resin is passing through what parts of
the fabric could be used to reduce the particle filtering, allowing for a more homogeneous
distribution of the strengthening particles.

Additionally, applying a detailed knowledge of the flow channels in the dual-scale fabric
could be used to create self-healing composites. Self-healing composites incorporate
particles, capsules, or tubes that are triggered to repair the part if it is damaged [42].
If the flow channels are post-filled with a solvable fluid, these could be used to create a
vascular network throughout the fabric. The vascular network could then be filled with
the healing agent.

1.4.2 Goals

Experiments carried out at the ETH CMAS Laboratory investigated the dual-scale
nature of resin flows in LCM processes [43]. Figure 1.7 shows how a second, colored
resin is injected behind a clear resin. The average velocities of the saturated (red) and
unsaturated (clear) flow fronts were compared, and these showed that there are regions
of the saturated flow that travel many times faster than the average flow front velocity.
The physical effect could be seen in the red, tree-like branches racing ahead of the rest
of the resin.

Such an effect shows that the fabric architecture creates a dual-scale flow. Darcy’s Law
can be used to accurately describe the flow in single-scale porous media. However, such
a model does not take into account that the saturated flow front has high variations in
Chapter 1. Motivation 13

Figure 1.7: Evident dual-scale flow in a two-color resin injection [43]

velocity. Volume averaging models have been applied to simulate the unsaturated flow
front in dual-scale textiles [24], but a model that considers the variations in velocity (i.e.
no averaging) for the saturated inter-tow flow does not yet exist.

The goal of this thesis is to create a robust framework that enables the simulation
and analysis of dual-scale flows for a wide variety of textile architectures. The results
developed here should aid in the development of a new mathematical model for dual-
scale saturated flows.
Chapter 2

Dual-Scale Textiles

Many fiber preforms used in LCM processes are dual-scale. Dual-scale fabrics exhibit
features on two distinct length scales: the individual fibers [µm], and the tows or yarns
[mm] into which the fibers are bundled together. These are the micro- and meso-scales
introduced in Section 1.3.1. The permeabilities in these two regimes can differ by orders
of magnitude, which gives rise to dual-scale flow [21].

Exceptions to dual-scale fabrics are random mats and unidirectional (UD) fabrics. In
random mats the individual fibers are not bundled together. And for UD fabrics, once
they have been processed the distinction between the bundles is no longer evident.

Woven, braided, knit, and 3D fabrics offer advantages in handling, drapability, and
impact resistance compared to UD or random mats. The material costs are also much
lower than their prepreg counterparts [36].

(a) Plain weave (b) Twill 2/2 (c) 3D fabric [44]

Figure 2.1: Different types of dual-scale textiles

14
Chapter 2: Material Characterization 15

Figure 2.2: Microscopy image of a woven aramid composite laminate. The light
circles are the individual fibers that make up the bundles which have been cut head on.
The dark lines are fibers that run along the direction of the cut.

2.1 Material Characterization

Even though composites are comprised of layers of a repeating fabric embedded in


a polymer matrix, the internal structure and distribution of fibers can vary. Fiber
bundles are spread out or sheared during layup, the layers are shifted and nested into
one another, and during resin injection individual fibers can be separated from their
tows. In contrast to the stochastic nature of a real composite part, a numerical model
based on repeating unit cells is in its essence periodic and repetitive.

In order to accurately model the flow of resin through fibers, realistic material param-
eters must accompany a precise numerical model. The complete information needed
to describe the geometrical fiber architecture is usually not available in manufacturer’s
data sheets and must be determined through tests.

This section describes what methods and data have been used to determine the geomet-
rical and material parameters used in the numerical flow model.

2.1.1 Geometrical Properties

The warp count, which specifies the number of warps and wefts per centimeter, is usually
given in the manufacturer’s data sheet. From this we obtain the yarn spacing. However,
to build the geometrical model of the fabric in TexGen, the yarn cross-sectional width
must be known. The difference between the width and spacing of the yarns results in the
inter-tow gap. This gap becomes a crucial factor in creating the flow channels through
the fabric, giving rise to dual-scale flow.
Chapter 2: Material Characterization 16

yarnspacing

yarnheight

yarngap
yarnwidth

Figure 2.3: Geometrical parameters needed to describe the yarns in a fabric

10 mm
yarnspacing =
warpcount (2.1)
yarnspacing = yarnwidth + yarngap

Cross-sections

The cross-sectional shape of the fiber bundles varies depending on the yarn width,
fabric compaction, and weaving pattern. The TexGen implementation [45] allows for
elliptical, lenticular, power-ellipse, hybrid, and polygonal cross-sectional shapes. By
using the power-ellipse function given in [37], the unit cell generator can model a wide
variety of fabrics by changing only a single parameter: n. Figure 2.4 shows different
yarn cross-sections that have been modelled using the power ellipse function (2.2).

h=2

n = 0:5
n = 1:0
0
n = 1:5
n = 2:0

!h=2
!w=2 !w=4 0 w=4 w=2
(a) Different exponents for the power ellipse function

(b) Implemented cross-sections for the dual-scale fabrics

Figure 2.4: Various cross-sections generated using the power ellipse function
Chapter 2: Material Characterization 17

w
C(t)x = cos(2πt) 0≤t≤1
 2
 h (sin(2πt))n

if 0 ≤ t < 0.5 (2.2)
2
C(t)y =
− h (− sin(2πt))n
 if 0.5 ≤ t ≤ 1
2

The specific cross-sectional shapes of the fabrics modelled in this thesis are determined
from microscopy images. For the plain weave fabric, a lenticular shape (n = 1.8) was
chosen on the basis of [37]. For the non-crimped biaxial fabric (n = 0.4), the shape was
based on the study given in [46].

Nesting

Nesting occurs between layers of a fabric because it is a lower energy state than perfectly
stacked layers. Nesting also occurs under compaction loads, where the individual layers
are pressed into one another. Potlouri and Sagar [47] modelled the compaction of textile
preforms and introduced a nesting factor (NF), which quantifies the amount of nesting
in a fabric:

ts
NF = Pn (2.3)
i=1 ti

where ts is the stack thickness, ti the individual layer thickness, and n the number of
layers in the stack. Initially there is no nesting: NF = 1. As the ratio of stack thickness
to total un-nested stack thickness decreases, we have stronger nesting (lower nesting
factor). Figure 2.5 shows the effect of compaction on the fabric nesting. The nesting
factor for the modelled plain weave unit cells was ≈ 0.80, for the biaxial ≈ 1.00.

Figure 2.5: Reduction in fabric height due to compaction pressure [47]


Chapter 2: Material Characterization 18

2.1.2 Dual-Scale Fabrics

The modelled glass fabrics are identical to those used in a concurrent thesis conducted
at the CMAS Laboratory, which by means of particle tracing, experimentally measures
the velocity distribution of resin flowing through dual-scale fabrics [48]. Using identical
input parameters allows for a tightly controlled comparison between experimental and
numerical results. This comparison is presented in Section 5.3.

The widths of both warp and weft yarns are measured using a WILD M3Z microscope.
Twenty measurements of warp and weft yarns throughout a fabric sample were made,
and the average widths are used.

Even though the plain weave and twill E-glass fabrics have identical areal weights and
use the same yarn types for both warp and weft yarns respectively, the microscopy
images in Figure 2.6 show significant differences in their fiber architecture.

Plain Weave

A plain weave fabric is a straightforward weaving pattern. The weft yarn passes above
and below each warp yarn in an alternating fashion. Both the warp and weft yarns
remain closely held together resulting in large gaps between fiber bundles (yarngap ≈
0.15 ∗ yarnspacing ).

The warp and weft label precisely identifies the glass fiber bundle. The “EC 9-204”
bundles are Electrical, Continuous, 9µm diameter glass filaments. The 204 refers to the
Tex number of the bundle, otherwise known as the yarn linear density, which is the
number of g/1000m of yarn.

The “EC 9-68x3 t0” is a non-standard name, which says something interesting about
the warp yarns. They have the same linear density as the weft yarns, (68x3 = 204), but
the warp bundles are comprised of three individual, never twisted glass rovings. The
“t0” specifies that they have zero twist. The three individual rovings can be seen upon
close inspection of Figure 2.6.

Tables 2.1 and 2.2 show the material and geometrical properties of the plain weave and
twill glass fabrics. The data was assembled from the manufacturer’s datasheets [49]
and microscope images. The standard deviation of the measured widths is shown in
parentheses.
Chapter 2: Material Characterization 19

(a) Plain weave (b) Twill 2/2

Figure 2.6: Close-up images of two types of glass fabrics. Warp yarns run vertically,
weft yarns horizontally.

Table 2.1: Plain weave glass fabric properties

Weave Pattern Areal weight [g/m2 ] Thickness [mm] Warps [p.cm] Wefts [p.cm]
Plain weave 280 0.30 7 6.5

Yarn Label Spacing [mm] Width [mm] Gap [mm]


Warp EC 9-68x3 t0 1.429 1.186 (0.040) 0.243
Weft EC 9-204 1.538 1.347 (0.052) 0.192

Table 2.2: Twill 2/2 glass fabric properties

Weave Pattern Areal weight [g/m2 ] Thickness [mm] Warps [p.cm] Wefts [p.cm]
Twill 2/2 280 0.35 7 6.5

Yarn Label Spacing [mm] Width [mm] Gap [mm]


Warp EC 9-68x3 t0 1.429 1.318 (0.033) 0.110
Weft EC 9-204 1.538 1.502 (0.044) 0.036

Twill 2/2

The Twill 2/2 weave pattern is a modification of the plain weave. The weft yarn goes
up over two and down under two warp yarns each cycle. Since the wavelength of the
twill fabric is twice that of the plain weave, the length of unrestricted yarn is also twice
as long. Because of this less stringent boundary condition, the twill yarns have more
freedom to move, as well as their cross-sections to flatten out. This results in twill
fabrics being well suited for draping over curved surfaces, leading to easier processing.

The over-under weaving pattern is offset by one yarn during each new pass of the shuttle
Chapter 2: Material Characterization 20

Figure 2.7: Close-up image of a UD carbon fabric. The Epikote finishing agent is
visible on the surface of the tow.

through the warp yarns, which creates the distinguishable diagonal “twill” pattern.
Compared to the plain weave, both the warps and wefts of the twill are more spread
out, leaving relatively narrow gaps in between the yarns (yarngap ≈ 0.05 ∗ yarnspacing ).

Non-Crimped Biaxial

Another type of reinforcing preform is the non-crimped fabric (NCF). Here the fiber
bundles are not woven together, they lie parallel to each other and are held together
with a much smaller stitching yarn. The parallel tows create a very uniform fabric
architecture, which allows for higher compaction, and therefore higher fiber volume
contents to be reached. NCFs can reach higher material properties than woven or twill
fabrics [50], but are more difficult to drape over complex (round) structures.

Figure 2.7 shows the Saertex carbon fabric modelled in this thesis, and Table 2.3 lists its
material properties. Some of the data comes from the manufacturer’s data sheet [51],
but the geometries of the yarns needed to be determined via microscopic measurements.

To create the 3D model of the biaxial NCF, two unidirectional (UD) layers are stacked
at 0°/90°. The tricot stitching that holds the bundles together will have an effect on the
formation of flow channels. However, because the stitching yarn is very small (6 g/m2 ),
the stitching is not considered in the numerical model.
Chapter 2: Material Characterization 21

Table 2.3: Carbon fiber UD fabric properties

Weave Pattern Areal weight [g/m2 ] Thickness [mm] Warps [p.cm] Wefts
Unidirectional 336 [–] 2.2 [–]

Yarn Label Spacing [mm] Width [mm] Gap [mm]


Warp T620SC 50C 24K 4.556 4.278 (0.058) 0.278

Additional Components
Stitching (Tricot) PES 76 dtex
Powder Momentive Epikote Resin 05390

2.1.3 Ethylcinnamate

The fluid modelled in the simulations is an Ethylcinnamate from Sigma Aldrich. It is


the same fluid used by the experiment described in Section 5.3. It was chosen by [48]
because its optical properties matched their experimental requirements. The refractive
index of Ethylcinnamate matches closely to that of the glass fibers described in Section
2.1.2. When the refractive index of both the fabric and the fluid are very similar, the
translucency of the fabric is greatly increased. This was essential for the particle tracking
algorithm, which tracks the particle-filled resin as it flows through the dual-scale fabric.

For the fluid dynamics simulation, the density ρ and dynamic viscosity µ must be known.
The fluid’s dynamic viscosity was not provided in the manufacturer’s data sheet and had
to be determined by a rheology test. The Ethylcinnamate was tested in a Mettler RM
180 Rheomat, run for 50s at 300Hz and at room temperature; the results are shown in
Figure 2.8. The dynamic viscosity for the fluid (at 23°C) was measured to be 82.3 mPa s.

It should be noted that this value is on the low range of viscosity for resins typically
used in LCM injections (100 - 500mPa s) [19]. Due to the low viscosity, the simulated
dual-scale flows will have slightly higher velocities than might be expected.

Table 2.4: Ethlycinnamate (CAS-Nr 103-36-6) material properties

Molar mass [kg/mol] ρ [kg/m3 ] µ [mPa s] refractive index


0.17621 1049 82.3 0.560
Chapter 2: Material Characterization 22

0.11 23
Viscosity
Avg =0.0823
Temperature
0.1

Temperature [/ C]
22.7
Viscosity [Pa"s]

0.09

22.4
0.08

0.07 22.1
10 15 20 25 30 35 40 45 50 55 60
Time [s]

Figure 2.8: Dynamic viscosity measurements of the Ethylcinnamate conducted at


room temperature
Chapter 2: 3D Modelling of Textiles 23

Figure 2.9: A 3D orthogonal fabric generated with TexGen

2.2 3D Modelling of Textiles

The 3D textile model lies at the basis of the dual-scale flow simulation. The geometrical
parameters that shape the model have a strong influence on the flow through the textile,
and must be determined carefully. This is why Section 2.1 closely investigated the glass
yarns and carbon fiber tows – the yarn geometries have a direct influence on the flow
channels created in the textile.

A textile consists of fiber bundles: when the fiber filaments are twisted together these
are called yarns, when they are untwisted and lie loosely side-by-side they are called
tows. Warp yarns span the length of the fabric, whereas weft yarns undulate through
the warps depending on the weave pattern. Small binder yarns can be used to hold the
fabrics together.

2.2.1 TexGen

Numerous textile generation software exist. WiseTex from K.U. Leuven [29] and TexGen
from the University of Notthingham [45] are perhaps the most popular. TexGen was
developed by Robitaille et al. in 1998 [52] and greatly extended by Sherburn in 2007 [37].
It is now a robust software, with ongoing development, comprehensive documentation,
and a wide user base in research.

We chose to use TexGen based on the following reasons: TexGen is free, open source, and
multi-platform. TexGen was designed with flexibility in mind: it can model practically
any type of fabric used in composite processing. Modelled textiles can easily be exported
Chapter 2: 3D Modelling of Textiles 24

into common CAD formats or directly as a mesh, which is useful for interfacing with
FE and CFD software. Finally, TexGen offers a Python scripting environment in which
the textile generation can be fully automated [44].

The general process of how TexGen builds a textile is as follows: first a yarn is defined
by master node positions. The nodes are connected with a Bezier or cubic spline, which
defines the yarn path. The yarn cross-section is extruded along this path, giving the yarn
its volumetric shape. Multiple yarns are then defined, woven through or around each
other, giving the fabric its distinctive weave pattern. The yarn volumes are meshed and
then restricted to a defined domain, resulting in an elementary unit cell which represents
the textile.

The textile generation process can be executed within the TexGen graphical user inter-
face (GUI), or programmed with the Python interface. In this thesis we made extensive
use of the Python interface, as it allowed for an automated generation of the different
unit cells needed in the CFD simulation.

The work of this thesis has been developed using TexGen 3.7.0, which was released on
7th August 2014. The compatibility of the code has also been tested with the latest
version: TexGen 3.8.0, released on 16th March 2015.

2.2.2 Generating a CFD Unit Cell

The code that was implemented to generate a repeating unit cell for the CFD simulation
can be broken down into four main steps: defining the textile properties, layering and
nesting, selecting a periodic domain, and exporting the 3D geometry. The detailed
process can be seen in Figure 2.10 and is described in the following sections. The
Python code is included in Appendix B.
Chapter 2: 3D Modelling of Textiles
Choose textile type: Layer textile: Select domain: Export unit cell:
(Plain weave, Biaxial, Twill)

Mat. data Add layers Trim domain STEP / IGES

Cross-section Nest layers FVC calculations ABAQUS mesh

Weave pattern yes


no
Interference? Export?

Generate yarns yes


no
Correct
interference
Refine weave

yes
Layers > 1? Build textile

no Import unit cell into


End
CFD simulation...

Figure 2.10: Flow chart of the unit cell generator implemented in Python/TexGen

25
Chapter 2: 3D Modelling of Textiles 26

Textile Properties

The unit cell generator requires the textile properties as an input. The fabric architec-
ture (plain weave, biaxial, or twill), yarn geometries, and material properties must be
specified by the user.

Even though the yarn width, gap size, and fabric thickness are known, the yarn thickness
is not. Without this value, the yarn cross-section cannot be defined. As a first approxi-
mation, both the warp and weft thickness are set to equal 1/2 the fabric thickness. As
is evident in Figure 2.6, the warp and weft widths can be different even though their
Tex numbers are equal. If this is the case, the Python script identifies the yarn with the
larger cross-sectional area, and then reduces its height until both warp and weft yarns
have equal area.

Using these geometrical properties, a CTextileWeave2D object is created. This object


holds the yarns and is the basic building block of the textile. Next the weave pattern is
defined. This is done by switching the over/under positions of the yarn paths. With the
cross-sections and weave pattern defined, the yarn paths and volumes can be generated.
At this stage the weave is checked for interfering points, which can occur when the yarns
are woven and their volumes intersect one another. Many problems arise when importing
CAD or mesh objects which contain intersecting volumes, so they need to be removed.
These interfering points are removed with the TexGen function RefineTextile() .

If this is to be a single-layered textile, the weave is built and meshed, then added and
stored by TexGen. If there are more layers, we proceed to the next step.

Layered Textiles

A layered textile consists of multiple weave objects stacked on top of each other. In
TexGen, are contained within a CTextileLayered object. The z-offset is calculated
from the height of each weave. At this stage, the layered textile looks like Figure 2.11b.
However, this perfect stacking of layers is not what woven fabrics actually look like in
an LCM process. In reality, the different layers nest into one another [47].

To achieve nesting in the 3D model, we first add an x- and y-offset to each layer. The x-
and y-offset is equal to 1/2 the weft and warp wavelength respectively. Then the function
NestLayers() is called. This function attempts to reduce the vertical spacing between
the individual layers without causing intersections. However, due to the wavy nature of
the fabric, intersections between the layers can still occur. If these are found they are
corrected by adding a small vertical offset. The finished layered textile is then built and
meshed, then added and stored by TexGen.
Chapter 2: 3D Modelling of Textiles 27

(a) Plain weave pattern (b) Layered textile

(c) Nested textile (d) Periodic unit cell

Figure 2.11: Stages for generating a periodic unit cell

Selecting a Periodic Domain

Choosing a periodic unit cell reduces the computational cost of the flow simulation.
Instead of calculating the flow through an entire textile, we can model only a single unit
cell. The domain is selected so that the geometry of the fabric is periodic on all faces.
This allows periodic boundary conditions to be set on the flow equations.

At least three layers are required to make a periodic (nested) unit cell: the middle layer
is nested from above and below. This ensures that there are no large flow channels
on the top and bottom layers, arise when a layer is not nested. This representation
simulates the flow in the middle of the fabric, not those layers that are adjacent to the
injection mold walls.

Modelling only a single unit cell is an idealization. Wong et al. [23] quantify the
variations to the local geometries that arise in real textiles. Gap sizes, yarn cross-
sections, and the amount of nesting are usually not uniform. They introduce statistical
variations to their RUCs, and build a representative elementary volume (REV) which
consists of multiple varying unit cell.

In this work, only a single repeating unit cell is chosen. The computational cost of
one RUC is already very high, and modelling multiple RUCs to make an representative
elementary volume is prohibitive.
Chapter 2: 3D Modelling of Textiles 28

Exporting the Unit Cell

The final step of the unit cell generator is to save the geometry in a standard file format
so that it can be imported into a fluid dynamics software. TexGen supports exporting
the CAD geometry to STEP and IGES file formats. The geometry, as well as material
properties (fiber orientation, stiffness, strength), can be exported as an ABAQUS input
file (.inp). This is especially useful if a mechanical analysis of the compaction of the
fabric is desired. TexGen can also export the geometry as a VTK unstructured mesh
or ABAQUS mesh. However, this export has a lesser level of control over the mesh size
and refinement than a dedicated meshing program.

For the integration into COMSOL Multiphysicsr , it was found that importing the STEP
geometry, and then meshing within COMSOL gave the most consistent and robust
results. This strategy was implemented for all the dual-scale flow simulations.

2.2.3 Resulting Textiles

Table 2.5 shows the unit cells which have been generated for the dual-scale flow simu-
lation. The same code was used to generate all the unit cells. In order to attain higher
fiber volume contents (Vf ), the fabric thickness (tf ab ) and thereby the yarn height was
reduced. The height reduction emulates a higher compaction pressure acting on the
fabric. This is based on the work of Chen & Chou [39], who showed that the height of
the woven fabric is reduced as the fiber volume fraction increases. The compaction also
results in an increase in the tow fiber volume content (Vf -tows ).

Table 2.5: Generated plain weave and biaxial unit cells

Plain weave glass Biaxial carbon


Vf Vf -tows tf ab [mm] NF Vf Vf -tows tf ab [mm] NF
0.331 0.421 0.470 0.800 0.440 0.532 0.867 1.00
0.350 0.447 0.443 0.800 0.480 0.581 0.795 1.00
0.371 0.472 0.418 0.799 0.502 0.609 0.760 1.00
0.382 0.487 0.405 0.799 0.520 0.631 0.734 1.00
0.401 0.512 0.385 0.800 0.540 0.655 0.707 1.00
0.420 0.540 0.366 0.802 0.560 0.680 0.682 1.00
0.441 0.568 0.348 0.803 0.580 0.706 0.658 1.00
0.600 0.731 0.636 1.00
Chapter 3

Dual-Scale Flow

The term dual-scale flow describes two very different types of flow occurring within
a fiber preform. This type of flow occurs when resin is injected into fabrics with a
dual-scale architecture. The two scales are the capillary channels between individual
fiber filaments and the gaps between fiber bundles (tows) [21]. The individual fibers
themselves are impermeable, but when bundled together in a tow, tiny spaces (∼µm)
between the fibers exist. This classifies the tows as a porous medium. Likewise, larger
spaces (up to 100 µm) between the different bundles form when they are woven into
a fabric. These inter-tow spaces become the free flow channels, which are orders of
magnitude wider than the inter-fiber spaces [26].

This chapter begins with the physics and equations used to model free and porous flows.
Various models for determining the permeability within the fiber bundles are compared.
Then the specific equations used to model the intra- and inter-tow flows are explained.

3.1 Porous and Free Flow

In order to describe a fluid’s flow through a porous medium, it is useful to define what
it means for a material to be porous. The Oxford Dictionary defines porous as having
minute spaces or holes through which liquid or air may pass [53]. The porosity is a
property that defines the ratio of empty space to the entire volume of a material. The
size of these minute spaces in relation to the liquid or gas determines if at all and with
how much effort a liquid or gas is able to permeate (pass through) the porous material.
The permeability measures the ease with which a substance can pass through the porous
medium – fluids require more energy to pass through materials with lower permeabilities.

29
Chapter 3: Darcy’s Law vs. Navier-Stokes 30

Figure 3.1: Darcy’s original sand column experiment [54]

3.1.1 Darcy’s Law

Henry Darcy (1803-1858) was a French engineer who made groundbreaking contributions
to the field of hydrology. In 1856 Darcy published his book, The public fountains of the
city of Dijon, in which he outlined the technical design and political implementation
of the new water infrastructure for the city of Dijon. He wrote, “A city that cares for
the interest of the poor class should not limit their water, just as daytime and light
are not limited” [55]. Buried within Appendix D of his manuscript, he presented an
experimentally derived formula relating the flow of water through a sand column, which
would later become known as Darcy’s Law:

s
q = k (h + e ± h0 ) (3.1)
e

He related the volumetric flow rate of water, q, to a permeability coefficient k and


geometry of the sand column (area over length: s/e), multiplied by the height of the
water on top of the column, h. The height of the water acted as the driving force pushing
the water through the porous sand column.

Today, Darcy’s Law is most often expressed as:

K
vdarcy = − ∇P (3.2)
µ
Chapter 3: Darcy’s Law vs. Navier-Stokes 31

which says that the average fluid velocity vdarcy is equal to the ratio of the material
permeability K and fluid (dynamic) viscosity µ, multiplied by the applied pressure
gradient ∇P . The negative sign arises due to fluid flowing from high to low pressures.

Darcy’s Law is relevant to many fields related to the flow of fluids in porous media
including: hydrology, hydro-geology, soil science, civil, petroleum, and composite en-
gineering [55]. Because the fibers in composite materials make up a porous medium,
Darcy’s Law became the de-facto equation used to calculate the resin flow during LCM
processes. Although Darcy’s Law is a simplification of the complex flow behaviour of
resin through the fiber bed, it has been used extensively to calculate flow patterns,
permeabilities, and filling times for composite parts [16, 19, 20, 22].

3.1.2 The Navier-Stokes Equations

In order to completely and precisely describe the flow of a fluid, we turn to the Navier-
Stokes equations. These powerful equations are applied in many fields to describe: the
aerodynamics of a plane, the flow of gas through a turbine, or the blood pumping from
one heart chamber to the next. Simply put, they are an expression of Newton’s Second
Law, applied specifically to the flow of fluids.

 
∂v
m = f
∂t |{z} (3.3)
| {z } External forces
Inertial terms

The Navier-Stokes equations are comprised of two parts: a conservation of momentum


and a continuity of mass equation.

 
∂v
ρ + v · ∇v = −∇P + µ∇2 v + f
∂t | {z } | {z } |{z} (3.4a)
| {z } Pressure gradient Viscous effects External forces
Inertial terms

∂ρ
+ ∇ · (ρv) = 0 (3.4b)
∂t

The conservation of momentum equation (3.4a) describes how changes to the fluid’s
velocity are driven by applied pressures, viscous effects, and external forces. The conti-
nuity of mass equation (3.4b) requires that the rates at which mass enters and exits a
system must be equal (i.e. mass must be conserved).
Chapter 3: Darcy’s Law vs. Navier-Stokes 32

Simplifying Assumptions

Solving the Navier-Stokes equations for a time-dependent, compressible flow in a com-


plex geometry is computationally expensive. In many circumstances, we can simplify the
full Navier-Stokes equations based on the assumptions of the physics being described.
For example, within the scope of this thesis, we consider our fluid (resin) to be incom-
pressible. That is, the density does not change with time ( ∂ρ
∂t = 0). Secondly, we consider
our flow to be steady. Steady flow (as opposed to transient flow) is valid for flows where
the flow field properties at a given location do not change with time ( ∂v
∂t = 0).

When considering incompressible, steady-state flow, the Navier-Stokes equations reduce


to:

ρv · ∇v = −∇P + µ∇2 v + f
(3.5)
∇·v =0

A final simplification of the flow equations can be made when the flow is very slow.
This is accurate for resin flow within a fiber tow, where the resistance to flow is high.
The Reynolds number is a dimensionless parameter which describes the ratio of inertial
forces to viscous forces in a flow [56], and is defined as:

ρvL
Re = (3.6)
µ

where L is the characteristic length scale of the flow problem (i.e. the diameter of a flow
channel). The intra-tow flows modelled in this thesis are very slow (µm/s), which leads
to Re  1. For such cases the inertial terms of the Navier-Stokes equations vanish,
giving Stokes flow:

0 = −∇P + µ∇2 v + f
(3.7)
∇·v =0

Applying the simplifications of this (slow) flow regime, the Navier-Stokes equations
become a linear system of differential equations, which reduces both the complexity and
computational effort required to solve them.
Chapter 3: Micro-scale Flow 33

Figure 3.2: Micro-scale model: an array of hexagonally packed cylinders

3.2 Micro-scale Flow

Micro-scale flow occurs within the fiber tows. We call it the “intra-tow” flow. It is the
flow of resin in the tiny spaces around the individual fibers. Because the spaces are very
small (usually below 10 µm), the resistance to flow is very high, which results in very
slow flow. This flow is orders of magnitude slower than in the inter-tow regions [21].

The micro-scale structure can be modelled as a porous medium, with the resistance to
flow characterized by its permeability. There are a number of semi-empirical formulas
for determining the permeability of the micro-scale structure of a fiber bed. To test the
validity of these models, a Navier-Stokes simulation has been conducted on the micro-
scale. The model which most closely matches the simulation has been implemented in
the meso-scale simulation.

3.2.1 Intra-tow Permeability Models

Semi-empirical formulas for the permeability within the fiber tows as a function of the
fiber volume content are well known, such as the Kozeny-Carman equation, or Gebart’s
formulas. Less common are the formulas from Berdichevsky & Cai, and Van Heusen &
Du Plessis. Senoguz [57] and Verleye [29] show that the simulated values for permeability
are in the same order of magnitude as these semi-empirical formulas. In order to find
the best model, we perform a similar verification in Section 3.2.2.
Chapter 3: Micro-scale Flow 34

Kozeny-Carman

In the field of composite materials, the Kozeny-Carman equation [58] is the most cited
equation for estimating the permeability of a fabric. Even though the formula was
derived for calculating the permeability of porous granular beds (not aligned fibers),
it is still commonly used as a rule of thumb for composite engineers [23, 29, 59]. The
Kozeny-Carman equation is:

r2 (1 − Vf )3
K= (3.8)
4kk Vf2

where kk is the Kozeny constant, and Vf the fiber volume content. The Kozeny-Carman
equations assumes isotropic permeability in the porous material, which is why it calcu-
lates only a single value for the permeability. It becomes clear that this formula was not
derived with fibers in mind, as the permeability parallel to the fibers varies greatly to
that in the perpendicular direction.

When the Kozeny-Carman equation is applied to porous textiles, the non-isotropic prop-
erties of the fabric can be incorporated into the Kozeny constant, kk . By adjusting the
values of kk , the equation can model Kk or K⊥ , but the constant must be determined
(experimentally or numerically) for each type of fabric.

Gebart

Gebart [34] derived permeability formulas specifically for fiber reinforced composites.
Taking the anisotropic nature of the fibers into account, he obtained two different formu-
las for the permeability parallel (along) and perpendicular to (transverse) the cylindrical
fibers:

8r2 (1 − Vf )3
Kk =
c Vf2
s 5 (3.9)
2
2 V m
K⊥ = r ∗ C1  − 1
Vf

where c is the shape factor1 , C1 is a packing constant, Vm the maximum fiber volume
content possible (for that packing), and r the cylinder radius. There is a strong similarity
between Gebart’s Kk and the Kozeny-Carman equation. They use slightly different
constants but the form of the equation is identical.
1
Verleye recommends changing Gebart’s constant c from 53 (originally proposed by Gebart) to 100
to attain a better fit for the longitudinal permeability [29].
Chapter 3: Micro-scale Flow 35

Table 3.1: Constants used for the permeability formulas with hexagonal packing

kk c C1 Vm c1 c2 c3 c4 m n
16 π
1.67 100 √
9π 6

2 3
0.111 -1.54 -2.82 0.229 -2.5 2.5

Berdichevsky and Cai

Berdichevsky and Cai [59] derive formulas for Kk and K⊥ based on their self-consistent
method. This method takes the heterogeneous fiber and empty space unit cell and
replaces it with a homogeneous porous medium. If the homogeneous, porous region
has a permeability K, which results in the same flow conditions (equal mass flow and
dissipation energy) as the Navier-Stokes flow within the heterogeneous unit cell, then
both regions are “self consistent”, and the permeability value of K holds in both regions.
This consideration results in:

" #
r2 1
Kk = ln 2 − (3 − Vf )(1 − Vf )
8Vf Vf
! (3.10)
r2 1 1 − Vf2
K⊥ = ln −
8Vf Vf 1 + Vf2

Cai and Berdichevsky

Cai and Berdichevsky [60] realized that the flow resistance varies with different packing
arrangements of the fibers: square, hexagonal, hollow square, and hollow hexagonal.
The subsequent derivations considered these differences and then they curve-fitted the
formulae to finite element results giving:

   
2 1 1 c2 Vf +c3 Vf2
Kk = r ∗ c1 ln e
Vm Vm
s m  s n (3.11)
Vf   Vf 
K⊥ = r2 ∗ c4  1−
Vm Vm

Van der Westhuizen & Du Plessis

Van der Westhuizen and Du Plessis [61] base their permeability formulas on a phase-
averaged formulation of the Navier-Stokes equation. They derived a closed form solution
Chapter 3: Micro-scale Flow 36

for unidirectional fiber beds. For the transverse permeability, they introduce an effective
fiber volume fraction (Vf∗ ), which takes into account the different resistance to flow in
hexagonal or square packing.

(5.299 − 2.157p )2p


Kk = r 2 ∗
48(1 − p )2
q (3.12)
(1 − Vf∗ )(1 − Vf∗ )2
K⊥ = r2 ∗
24(Vf∗ )1.5

3.2.2 Intra-tow Permeability Simulation

In order to calculate the permeability, two simulations have been run: one for parallel
flow along the fibers, and the other for the perpendicular flow through the fibers. The
simulation meshes and flow fields are shown in Figure 3.3, and the simulation parameters
are listed in Table 3.2.

Both simulations implement incompressible, steady-state flow (3.5) on a parallel array of


hexagonally packed cylinders. The hexagonal packing was chosen because it allows for a
higher maximum fiber volume content (Vm ≈ 0.907). Both the axial and transverse flow
simulations use rectangular unit cells. No-slip boundary conditions (u = 0) are applied
to the cylinder walls, whereas slip boundary conditions (u · n = 0) are applied to the
unit cell walls.

The primary difference between the Kk and K⊥ simulation is the direction of the pressure
gradient. For Kk , the pressure gradient drives the flow axially along the cylinders: for
K⊥ , perpendicularly between the cylinders.

The permeability is calculated from the velocity field at the mid-plane of each unit
cell. To calculate K for an increasing range of Vf , the separation between two adjacent
cylinders is reduced. This is controlled by setting the distance between the center points
of adjacent cylinders. When Vf ≥ Vm , the cylinders intersect. For fibers (cylinders) with
hexagonal packing the separation is:


πr2
d(r, Vf ) = √ (3.13)
3
2Vf

The geometry is re-meshed for each new iteration of Vf , with the mesh size also depen-
dent on the cylinder separation distance d(r, Vf ). This ensures that the element size
does not restrict the flow in the narrow regions between the cylinders, even if the fiber
volume content is high.
Chapter 3: Micro-scale Flow 37

(a) Kk meshed unit cell (b) Kk flow field

(c) K⊥ meshed unit cell (d) K⊥ flow field

Figure 3.3: Axial and tranverse permeability simulations (Vf = 0.60)

Table 3.2: Micro-scale simulation parameters

Kk and K⊥ simulation parameters


Flow equations Navier-Stokes (incompressible, steady-state)
Packing hexagonal
Fiber radius 4.5 µm
Unit cell dimensions change as a function of Vf
Cylinder BC no-slip (u = 0)
Wall BC slip (u · n = 0)
∆P 1000 Pa
Element type P2-P1 tetrahedral
Element size (boundary) 0.40 ↔ 3.20 % d(r, Vf )
Element size (domain) 1.25 ↔ 6.25 % d(r, Vf )
Chapter 3: Micro-scale Flow 38

The permeability of the micro-scale unit cell can be extracted from the simulated velocity
field by means of Darcy’s Law. The velocity field is averaged over the cross-section of
the flow channel. Only the component of the velocity vector pointing in the direction of
the permeability to be calculated is used. Darcy’s law considers the flow in the entire
porous medium (fibers and flow channels), consequently the average velocity vavg must
be multiplied by the porosity to give the Darcy velocity vdarcy . Then the permeability
of each flow direction is calculated using Darcy’s Law:
R
ARv(x) dA
vavg =
A dA

vdarcy = vavg p (3.14)

µ µ
Kk = −vdarcy (xi ) and K⊥ = −vdarcy (xj )
∇P ∇P

3.2.3 Results

Figures 3.4 and 3.5 compare the longitudinal and transverse permeability values of the
semi-empirical models to those obtained from the Navier-Stokes flow simulation.

We immediately see that the Kozeny-Carman equation (3.8) overestimates the perme-
ability for both the longitudinal and transverse flow. Even though kk = 1.66 is the value
given for a hexagonal array of cylinders [29], the constant would need to be fit to the
simulation results. The weakness of the Kozeny-Carman equation is that its accuracy
depends on the constant kk being experimentally or numerically fit to the fabric being
modelled.

For Kk , we see that the numerical values are in good agreement with the values obtained
from Gebart (3.9) and Berdichevsky-Cai (3.10). Westhuizen & Du Plessis (3.12) over-
estimates Kk . The best correlation is with Cai-Berdichevsky (3.11), which is accurate
even at the highest range of Vf .

For K⊥ , Gebart and Cai-Berdichevsky give almost identical results. This is to be ex-
pected, as these two equations (and their constants) are very similar. For higher fiber
volume contents, the simulated permeability begins to decrease. Westhuizen & Du
Plessis’ model follows this trend up until Vf = 0.50, but then overestimates the drop.

On the basis of this comparison, we have chosen to use the improved model of Cai
& Berdichevsky (3.11) to calculate the intra-tow permeability. By encapsulating the
micro-scale fiber architecture into a representative permeability value, we can treat the
meso-scale tows as a single porous medium.
Chapter 3: Micro-scale Flow 39

Kozeny & Carman (1937)


Gebart (modi-ed by Verleye 2008)
Berdichevsky & Cai (1993)
10!11 Cai & Berdichevsky (1993)
Westhuizen & Du Plessis (1996)
NS Simulation
Permeability [m2 ]

10!12

10!13

0.2 0.3 0.4 0.5 0.6 0.7 0.8


Fiber volume content [-]

Figure 3.4: Permeability Kk for increasing Vf - Longitudinal Flow

10!10
Kozeny & Carman (1937)
Gebart (1992)
Berdichevsky & Cai (1993)
10!11 Cai & Berdichevsky (1993)
Westhuizen & Du Plessis (1996)
NS Simulation
Permeability [m2 ]

10!12

10!13

10!14

10!15
0.2 0.3 0.4 0.5 0.6 0.7 0.8
Fiber volume content [-]

Figure 3.5: Permeability K⊥ for increasing Vf - Transverse Flow. Because the models
of Gebart and Cai & Berdichevsky are almost identical, their lines overlap.
Chapter 3. Meso-scale Flow 40

Intra-tow
Inter-tow

Figure 3.6: Intra- and inter-tow flow regions in a meso-scale unit cell

3.3 Meso-scale Flow

Meso-scale flow occurs between the fiber tows. We call it the “inter-tow” flow. It
describes the flow of resin through the gaps between neighboring tows. As introduced
at the beginning of Chapter 3, the size of the flow channels is orders of magnitude larger
than the spaces between individual fibers. This results in two different flow regimes, a
dual-scale flow.

Modelling the meso-scale flow through a fabric while still resolving the individual fibers
is computationally prohibitive. This is why the problem is split into a intra- and inter-
tow (micro- and meso-scale) simulation. The many individual fibers in a tow are ho-
mogenized into a single porous medium, with the representative intra-tow permeability
encapsulating the micro-scale resistance to the flow.

The goal of this thesis is to investigate the velocity distribution of the meso-scale flow.
With this in mind, we do not use a volume averaging technique to calculate the meso-
scale flow, which would loose (average away) the details of the flow field. To capture
the full flow field, the Navier-Stokes equations are used.

In the intra- and inter-tow regions, we use different forms of the Navier-Stokes equations
that are specific to the flow physics in that region. For both cases, we can apply these
simplifications to the Navier-Stokes equations introduced in Section 3.1.2:

∂u
Steady-state flow: ∂t =0

∂ρ
Incompressible fluid: ∇ρ = 0, ∂t =0
Chapter 3. Meso-scale Flow 41

3.3.1 Intra-tow flow

In the intra-tow region, where the flow is very slow, we use the Stokes-Brinkman equa-
tions [62]. The Brinkman formulation extends the Stokes equations (3.7) with a perme-
ability term K, which is needed because the tows are a porous medium.

" #  
µ T
 µ
0 = ∇ · −pI + ∇u + (∇u) − u+f
p K (3.15)
ρ∇ · u = 0

On close inspection, we see that the Brinkman formulation adds a term that equals the
pressure drop in a porous medium described by Darcy’s Law (3.2). Additionally, the
viscous term of the Stokes equation is divided by the porosity p . This represents the
effective viscosity (µef f ≈ µ/p ) in the porous region, which has been analytically and
numerically shown to be larger than the free-flow viscosity [63].

The values of the permeability tensor K are determined using the model of Cai &
Berdichevsky (3.11). In this way, the Stokes-Brinkman equations couple the micro- and
meso-scale flow models.

3.3.2 Inter-tow flow

The incompressible, steady-state Navier-Stokes equations [62] are well suited to describe
the flow in the inter-tow channels.

  
T
ρ (u · ∇u) = ∇ · −pI + µ ∇u + (∇u) +f
(3.16)
ρ∇ · u = 0

In this region, we cannot assume that the flow is very slow. On the contrary, the
simulations will show that the flow speeds in this region can be quite fast. Therefore,
the inertial terms of the Navier-Stokes equations (ρ (u · ∇u)) are kept, resulting in a
non-linear system of differential equations.

The same pressure and velocity fields (p and u) are solved for in both intra- and inter-
tow regions. This ensures continuity of both the pressure field and fluid velocity at the
Chapter 3. Meso-scale Flow 42

boundary between the two regions. More details on the interaction between the two
flow equations are given in [62].
Chapter 4

Solving the Dual-Scale Flow

Analyzing the velocity distribution of resin flowing through a dual-scale textile requires
a multi-step solution process. The first step is generating a periodic unit cell, which has
been described in detail in Section 2.2.2. With the geometrical model ready, the fluid
dynamics simulation can be set up. Here the boundary conditions, flow equations, and
solver strategy are defined. Then the simulation must be solved, and finally the results
extracted and analyzed.

For each stage in the solution, a different environment is used. This chapter describes
the second and third steps (simulation and computing), and how they are integrated to
form a complete analysis framework.

COMSOL
(simulation)
MATLAB
TexGen EULER HPC
• import (analysis)
(geometry) (computing)
geometry
• assemble
• material • set flow eqs. • FTP files
data
props. • set BCs • set up
• velocity
• fabric type • mesh modules
distributions
• generate geometry • submit job
• flow rates
unit cell • define solver • analyze log
• plotting
• define data
export

Figure 4.1: Flow chart for the numerical analysis of dual-scale velocity distributions

43
Chapter 4: COMSOL Model 44

4.1 COMSOL Model

In this thesis, COMSOL Multiphysicsr (version 4.4) was used to simulate the dual-scale
fluid dynamics problem. As its name implies, COMSOL Multiphysics is a simulation
suite specialized in coupling multiple physical domains. It has implementations for
structural mechanics, heat transfer, fluid dynamics, electrodynamics and more [64].
The user can, in a relatively straightforward way, set up the coupling equations to
model interactions between the domains.

The dual-scale flow problem has been implemented using the Free and Porous Media
Flow interface, which is part of the Porous Media and Subsurface Flow branch within
COMSOL’s CFD Module [62].

4.1.1 Geometry Import and Input Parameters

The first element needed for the COMSOL simulation is the unit cell geometry. The file
that was generated by the unit cell generator (see Section 2.2.2) is imported into COM-
SOL using the CAD Import Module. This module enables the import of most common
CAD file types. The import of the unit cell will fail if the geometry has any intersecting
volumes, which is why the interference detection and removal was implemented in the
unit cell generator code.

Because a unique unit cell is used for each fiber volume content, the simulation needs to
be rebuilt every time. In order to automate the process of setting up the flow equations,
applying boundary conditions, and meshing the intra- and inter-tow domains, these
processes were parametrized according to the unit cell geometry. The unit cell width,
height, length, as well as other fabric properties, are saved as variables. The selec-
tion of the different domains (intra- or inter-tow flow, periodic boundaries, inlet/outlet
boundaries), is executed on the basis of the geometrical parameters of the unit cell.

4.1.2 Materials

As we are modelling fully-saturated flow, the entire unit cell is filled with a fluid. Ethyl-
cinnamate is implemented as the fluid. This is the same fluid used in the experimental
characterization of the dual-scale velocity distribution [48], and its properties are found
in Section 2.1.3.

For the fabric, a distinction in the intra-tow region is made between the parallel and
perpendicular tows. This refers to the orientation of the fibers relative to the direction
Chapter 4: COMSOL Model 45

(a) Inlet and outlet (b) Periodic boundaries

(c) Parallel (gold) and perpendicular (d) Inter-tow domain


(red) tows

Figure 4.2: Automated selections for various model parameters

Table 4.1: Meso-scale simulation parameters

Physical input parameters


pin 200 Pa inlet pressure
pout 0 Pa outlet pressure
r 4.5 µm fiber radius (glass)
π
Vm √
2 3
maximum FVC (hexagonal)
Vf -tows (value from TexGen) intra-tow FVC

Geometrical input parameters


warpSpacing 10 mm / warpcount spacing between warps
weftSpacing 10 mm / weftcount spacing between wefts
nWarps 2.0 number of warps
nWefts 2.0 number of wefts
ucTol 0.01 mm tolerance used for selection
ucX0, ucY0, ucZ0 (values from TexGen) initial coords. of unit cell
ucL, ucW, ucH (values from TexGen) length, width, height of unit cell

Simulation parameters
Element type P2-P1 tetrahedral quadratic velocity, linear pressure
Number of DOF ≈ 1.5 million varies slightly for each unit cell
Solution type stationary incompressible, steady-state flow
Solver MUMPS direct and distributed solver
Relative tolerance 0.001 terminates when > relative error
Residual factor 1000 used for termination criteria
Chapter 4: COMSOL Model 46

of the flow. The flow direction has been chosen to be in the x̂ direction. This means that
all warp tows have their fibers aligned parallel to the flow direction, and all weft tows
perpendicular to it. The orientation of the fibers has an influence on the micro-scale
permeability of the tows.

The values for Kk and K⊥ are calculated directly from the model of Cai & Berdichevsky
(3.11), and depend on the tow fiber volume content (Vf -tows ) of the unit cell under study.
The permeability tensors for the parallel and perpendicular tows are both symmetric,
and contain entries only along the diagonal. They differ in the placement of the Kk
term: for the parallel tows the axial permeability is in the x̂ direction, and for the
perpendicular tows it is in the ŷ direction. The change in permeability along the yarn
path due to their waviness is not considered.

 
K
 k 0 0 
Kparallel-tows =  0 K⊥ 0 
 
 
0 0 K⊥
  (4.1)
K⊥ 0 0 
Kperp-tows =  0 Kk 0 
 
 
0 0 K⊥

4.1.3 Flow Equations and Element Order

The Free and Porous Media Flow interface applies different flow equations to the porous
and non-porous (free) domains of the model. The intra-tow domains (belonging to the
yarns) are solved with the Stokes-Brinkman equations (3.15). The inter-tow domains
are solved with the Navier-Stokes equations (3.16). For a detailed explanation of these
equations, see Section 3.3.

The driving force for the flow through the unit cell is a pressure gradient applied in the
x̂ direction. The pressure drop over the unit cell was determined by the experiment.
The experiment applied a pressure difference of 15.0 kPa from pot to pot. The resulting
pressure drop over the dual-scale fabric was 11.5 kPa. Dividing this by the length of the
entire fabric (180.0 mm) results in ∆p ≈ 200 Pa for a 3.08 mm long unit cell.

A detailed explanation of how the pressure gradient over the unit cell is calculated can
be found in Appendix C.
Chapter 4: COMSOL Model 47

Figure 4.3: Flow-driving pressure drop (Pa) over unit cell

In COMSOL, the element order is set on the physics interface level. Within the CFD
module, the discretization can be set individually for the solution of the pressure and
velocity fields. Linear elements (P1) are chosen to model the pressure field, as its gradient
is smooth over the length of the unit cell. However, due to the dual-scale nature of the
fabric, there are large gradients in the velocity field. Linear elements would not be
able to accurately describe the large changes in velocity, especially in the flow channels.
Second order elements (P2) are implemented to solve the velocity field, even though
their numerical cost is higher. This discretization is called P2-P1 (O2 (v) − O1 (p)).

4.1.4 Boundary Conditions

Both the geometry and the boundary conditions of the unit cell are periodic. Figure 4.4
shows how periodic flow conditions are applied to the top & bottom, and left & right
walls of the unit cell. Periodic flow conditions apply the following constraint equations
to the selected walls:

usource = udest
(4.2)
psource = pdest

By implementing periodic boundary conditions, the results calculated from a single unit
cell can be applied to the flow within an entire fabric. Granted, a single repeating
unit cell (RUC) represents an idealized fabric and does not take into account occurring
variations. It does however allow for the simulation to resolve a level of detail in the
dual-scale flow that cannot be computed in a larger model.
Chapter 4: COMSOL Model 48

Periodic Boundaries
(top & bottom)

Periodic Boundaries
(left & right)

Flow direction
(u)

Figure 4.4: Boundary conditions applied to top & bottom and left & right walls

4.1.5 Solver Settings

The meso-scale simulation is conducted as a stationary study. This type of study solves
equations without any time derivatives, and is based on the steady-state simplification
of the Navier-Stokes equations. The nonlinear solver applies a fully coupled, damped
Newton method with an initial damping factor of 0.01. COMSOL’s nonlinear solver
algorithm automatically adjusts the damping factor until the solution converges [62].

The termination criteria is based on the residual tolerance. The solution’s relative error
is calculated from the residual factor times the residual-based error. The nonlinear
iterations terminate when the relative tolerance is greater than the relative error (0.001).
For more details, see Termination Criterion in [64].

Initial attempts at solving the full model, which contained over a million degrees of
freedom (DOF), did not converge. The solver was changed from the GMRES iterative
to the MUMPS direct solver, and convergence was achieved. Direct solvers are generally
very robust, however they require more memory. The memory usage scales with the
number of DOF1.5 to DOF2.0 [62], which is why the memory usage of the flow simulation
was very high. Usage of the ETHZ EULER cluster enabled big memory simulations,
and because MUMPS is a distributed solver, it was well suited for cluster computing.
Chapter 4: Meshing 49

Figure 4.5: The mesh of the dual-scale unit cell

“Numerical simulations of resin flow in the complex shaped tow need heavy
computations with a huge number of elements, since the tow size is so small
and very tiny elements are required.” [22]

4.2 Meshing

The geometry of a plain weave fabric (a dual-scale fabric) is not ideal for meshing.
Figure 4.5 shows the sharp corners of the lenticular cross-sections of the yarns, and
that the small spaces between neighboring yarns require a high resolution mesh. These
factors result in a numerical simulation with many elements and accompanying degrees
of freedom.

The flow channels between the tows exhibit large gradients in the velocity field. In order
to resolve such large changes in the velocity, enough elements are needed in the inter-
tow region. If there are too few elements, the flow is artificially restricted by the poor
mesh. A convergence study has been performed to determine the appropriate number
of elements required to accurately solve the dual-scale flow model.

4.2.1 Convergence Study

Preliminary simulations confirmed that the intra-tow flow is very slow and also has small
velocity gradients. For this reason more weight was put on finely meshing the inter-tow
regions, where the velocity gradients are large.
Chapter 4: Meshing 50

(a) Inadequate mesh with choked velocity field

(b) Mesh with 2 boundary layers - good flow field but high computational cost

(c) High resolution tetrahedral mesh - good flow field at moderate computational cost

Figure 4.6: Comparison of three meshing strategies

Different meshing approaches for the inter-tow region have been evaluated. The first
approach simply increases the number of tetrahedral elements in the inter-tow domain
by reducing the maximum element size. The element sizes are given as a percentage of
the cross-section diagonal length.

A course mesh (max el. size > 0.51%) could not resolve the geometry of the unit cell,
and would cause the meshing algorithm to fail. Maximum element sizes from 0.51% to
0.17% were sampled, producing a range of 0.8 to 1.6 million degrees of freedom (DOF)
for a single unit cell.

The addition of more elements, particularly in the flow channel regions, resulted in an
improved flow field. Such a flow follows one of Navier-Stoke’s fundamental principles
of fluid velocity: the farther the fluid is from the contact surfaces, the faster it flows.
The choked flow channels result in artificially slow regions which are inconsistent with
Navier-Stokes. The difference between the central flow channel of Figures 4.6a and 4.6c
shows this improvement.
Chapter 4: Meshing 51

0.144 33:36
Only TETS
With BND1
0.141 With BND2 28:48
Computational time
Average inter-tow velocity [mm/s]

0.138 24:00

Computational time [hrs]


0.135 19:12

0.132 14:24

0.129 09:36

0.126 04:48

0.123 00:00
0.8 1 1.2 1.4 1.6 1.8
Degrees of freedom (millions)

Figure 4.7: Convergence study for plain weave fabric (Vf = 0.33, ∆p = 200 Pa)

The second approach (Fig. 4.6b) was to add a boundary layer on the intra- inter-tow
interface. The boundary layer extends into the inter-tow domain, and was tested with a
thickness of 1 and 2 brick elements. The additional elements improved the flow field, but
significantly increased the computational cost. This is because COMSOL’s quadratic
brick elements have 8.5 nodes per element, whereas the quadratic tetrahedral elements
have only 1.4 nodes per element [65]. More nodes per element result in more degrees of
freedom.

Figure 4.7 shows the results from the convergence study. The average inter-tow velocity
of the cross-section is used as a scalar parameter for the comparison. Three meshing
strategies are shown consisting of: only tetrahedral elements, and tetrahedrals combined
with 1 and 2 brick elments in the boundary layer.

The meshes with boundary layers have many DOF, even though they have less elements
than the tetrahedral only mesh. To keep the same number of DOF, adding the boundary
layer comes at the cost of reducing the number of tetrahedral elements in the inner inter-
tow region. Additionally, because the boundary layer uses quadratic brick elements,
the computational time increases much more dramatically than the mesh with only
tetrahedral elements. There the computational time increases linearly with the number
of DOF, which is a significant advantage.
Chapter 4: Meshing 52

The tetrahedral mesh shows two convergence plateaus. The first occurs at a low number
of DOF, where elements of the coarse mesh are not able to fully map the high gradients
of the Navier-Stokes flow. This results in a lowering of the average velocity, as the
high flow regions cannot fully develop. Increasing the number of DOF above 1.2 million
brings the simulation to the second convergence plateau.

The convergence study has given valuable insights for the optimal mesh parameters.
For the final simulations, a tetrahedral only mesh with at least 1.5 million DOF was
implemented. Such a mesh is able to model the fully developed flow in the inter-tow
channels, is sure to be on the second convergence plateau, and has a linear dependence
between the CPU time and number of DOF.
Chapter 4: EULER High Performance Cluster 53

4.3 EULER High Performance Cluster

The convergence study shows that in order to simulate a fully developed flow field, the
meso-scale simulation should have around 1.5 million DOF. These simulations require
high computational effort, and have been executed on the ETH EULER cluster.

EULER is an abbreviation for Erweiterbarer, Umweltfreundlicher, Leistungsfähiger ETH-


Rechner. Translated this means expandable, eco-friendly, powerful, ETH-computer. Its
name is also an homage to the 18th Century Swiss mathematician and physicist Leon-
hard Euler, who formulated one of “the most beautiful mathematical formula ever”
[66]:

eiπ + 1 = 0

The EULER cluster became operational in 2014, extending the capabilities of the ETH
BRUTUS cluster. The EULER was designed for speed and contains 416 compute nodes,
each fitted with: two 12-core 2.7GHz Intel Xeon processors with 64GB (up to 256GB)
1866 MHz RAM [67].

4.3.1 Running COMSOL Simulations on EULER

Detailed instructions for running COMSOL Multiphysics simulations on the BRUTUS


and EULER clusters can be found on the cluster wiki page [68]. The basic steps are
outlined as follows:

• Prepare simulation on local PC

• Move files onto EULER filesystem using FTP client

• Establish an SSH connection with usrname@euler.ethz.ch

• Load COMSOL module into user profile

• Determine parameters for batch submission

• Submit batch job to LSF queue (and wait...)

• Analyze log and retrieve completed simulation file


Chapter 4: EULER High Performance Cluster 54

The essential commands for navigating the CentOS 6.5 operating system on EULER,
and for submitting jobs to the LSF batch system are given in the following code block.

# Load the environment for COMSOL 4.4


module load comsol /4.4
# View your current usage/jobs
busers
bjobs

# Example LSF batch submission


bsub −n 6 −W 8:00 −R " rusage [mem =30000]" comsol batch −clustersimple −
inputfile PW FVC 042 .mph −outputfile P W F V C 0 4 2 n 6 8 h r 3 0 G B .mph

The LSF parameters −n 6 −W 8:00 −R "rusage[mem=30000]" specifies that 6 cores with


30GB of memory allocated for each should be used. The job will be killed if it exceeds
the 8:00hr time limit.

A typical flow simulation for the plain weave fabric running on EULER with 6 cores
required between 3 to 4hrs to solve.

Table 4.2 shows a summary of the computational resources used during this thesis.
Initially test simulations were run on BRUTUS, and the final dual-scale models were
solved using EULER. At this point it is appropriate to express my sincere thanks to
IT Services for providing student access to the exceptional computing facilities at ETH
Zürich. Without this resource, the scientific research undertaken in this thesis would
not have been possible.

Table 4.2: Computational workload on BRUTUS and EULER clusters

Number of jobs CPU time [hrs] [days]


BRUTUS 143 2450.8 102.12
EULER 67 865.2 24.61

Total 210 3316.0 138.17


Chapter 4: Data Extraction 55

Figure 4.8: A zoomed-in view of the regular grid which samples the cross-sectional
velocity field. This large flow channel, which has high velocity gradients, is sampled by
over 700 data points.

4.4 Data Extraction

The dual-scale simulation enables a numerical analysis of the velocity distribution of the
resin in the textile. In order to extract the velocity data, a regular grid is laid over a
cross-section of the unit cell. The x̂-component of the velocity field at each point in the
regular grid is evaluated on the quadratic shape functions of the CFD solution.

The cross-section is a ŷ-ẑ cut-plane that is perpendicular to the direction of flow (x̂)
and has been chosen to be at 3/4 length of the unit cell. This far into the unit cell
the flow is well developed. Changing the location of the cut plane (as long as it is not
near the inlet or outlet) does not have a significant effect on the shape of the velocity
distribution.

The COMSOL regular grid samples the velocity field within the cross-section, not in-
cluding the boundaries. As its name implies, the points of the regular grid are spaced
evenly over the entire cross-section. This means that, per unit cell, each data point
covers an equal amount of area.

The dimensions of the regular grids used to sample the plain weave and biaxial simu-
lations are listed in Table 4.3. Because the cross-sectional area of the biaxial unit cells
are approximately four times larger than those of the plain weave, four times as many
points are used.

For each fabric architecture, multiple simulations are carried out spanning a range of
fiber volume contents. As the Vf is increased, the thickness of the unit cell is reduced.
It was determined that the total number of data points sampled by the regular grid
should remain the same for the different Vf simulations. This allows for a more uniform
Chapter 4: Data Extraction 56

Table 4.3: Dimensions for the plain weave and biaxial regular grids

Grid (m × n) Total points


Plain weave 400 × 80 32000
Biaxial 600 × 200 120000

comparison of the velocity distribution histograms. Due to the change in height of the
unit cell, there is a corresponding change in the area sampled by the data points.

These considerations affect only the resolution of the data extracted for the velocity
histograms and volumetric flow rate calculations. They do not determine the resolution
of the CFD flow simulation: that is given by the mesh density and element order.
Chapter 5

Results and Discussion

This chapter presents the results obtained from the meso-scale CFD simulations. First,
a detailed analysis of the velocity distribution of flow through the dual-scale unit cell
is made. Then a comparison between two dual-scale fabrics, a plain weave and biaxial,
is presented. Another comparison is made to the velocity distributions found in an
experimental investigation. Finally, the results presented in this thesis are shown to
expand on those currently available in state of the art dual-scale flow research.

5.1 Velocity Distribution in a Plain Weave

An analysis of the velocity distribution gives additional insight into the resin flows in the
dual-scale fabric. From it we can understand the spatial influence of the flow channels
on creating regions of fast flow and transporting resin through the fabric.

Figure 5.1: Resulting flow field of the dual-scale simulation

57
Chapter 5: Velocity Distributions 58

600
Number of data points [-] 500 Inter-tow velocity (n = 8768)

400
300
200
100
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35
Velocity [mm/s]
Number of data points [-]

8000 Intra-tow velocity (n = 23232)

6000

4000

2000

0
0 1 2 3 4 5
Velocity [mm/s] #10!3

Figure 5.2: Velocity distribution for plain weave fabric (Vf = 0.40)

5.1.1 Histograms

The most straightforward method of analyzing the velocity distribution is to make a


histogram of all the velocity data points. Such a histogram shows the distribution of the
fluid velocity within the textile. Due to the dual-scale nature of the flow, the velocity
data points are split into two regimes. This ensures that the intra-tow regime, which
has many more data points, does not obfuscate the distribution of the inter-tow flow.
Both distributions are shown in Figure 5.2.

The dataset for the histograms consists of the x̂-component of the velocity field u. This
is the direction of the flow along the length of the unit cell. Because the intra-tow flow
is solved as a flow through a porous medium, the actual fluid velocity is larger by a
factor of the tow porosity (1 − Vf -tows ). As there are no fibers present in the inter-tow
region, the velocity remains the same.

uintra
vintra =
1 − Vf -tows (5.1)
vinter = uinter
Chapter 5: Velocity Distributions 59

The histograms are calculated using the MATLAB function histogram() , and split
the data into 50 bins. With nbins = 50, individual features of the velocity distribution
can be resolved, without introducing noise.

Figure 5.2 shows two interesting phenomena. First, the simulation confirms that the
meso-scale flow is approximately three orders of magnitude faster than micro-scale flow.
This is expected because of the the low permeability of the tows. Second, the number
of data points in the inter-tow region is much lower than the intra-tow region. A large
fraction of the total cross-sectional area is occupied by the tows.

A small number (< 1%) of the intra-tow data points shown in Figure 5.2 have negative
velocities. This is likely due to numerical error: when the relative tolerance is reached
and the simulation is terminated, a few small areas of the slow intra-tow flow (µm/s)
have not yet reached a stable solution.
Chapter 5: Velocity Distributions 60

100

Vf = 0:33
10!1 Vf = 0:35
Area [mm2 ]

Vf = 0:37
10!2 Vf = 0:38
Vf = 0:40
Vf = 0:42
10!3 Vf = 0:44

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4


Velocity [mm/s]

100

Vf = 0:33
!1 Vf = 0:35
10
Area [mm2 ]

Vf = 0:37
Vf = 0:38
10!2 Vf = 0:40
Vf = 0:42
Vf = 0:44
10!3

0 2 4 6 8 10 12
Velocity / Vavg [-]

Figure 5.3: Distribution of the cross-sectional area covered by increasing velocity


regimes. Both the absolute and normalized velocities of the full data set (intra & inter)
are sampled.

5.1.2 Effect of Fiber Volume Content

Higher Vf result from an increased compaction of the fabric, which reduces both the
micro- and meso-scale spaces in the fabric [47]. This in turn will restrict the flow, giving
lower velocities in the higher Vf fabrics. This phenomenon is clearly shown in Figure
5.3, where the maximum velocity of Vf = 0.33 is almost twice as fast as Vf = 0.44.

When comparing multiple simulations with different velocities, it makes sense to nor-
malize the velocities by the average velocity. This allows for a comparison of the shape
of the velocity distribution, even though the absolute velocities have different ranges.

In an effort to make the velocity distributions more readable, the histogram bar plots
have been changed to line plots. Also, we change the y-axis from the number of data
points to the area covered by these points, as this is a physical quantity related to the
flow field.

The line histograms are implemented with the MATLAB function histcounts() .
This function returns the number of points for each binned velocity value. Multiplying
Chapter 5: Velocity Distributions 61

0.5
Vf = 0:33
0.4 Vf = 0:35
Vf = 0:37
Area [mm2 ]

0.3 Vf = 0:38
Vf = 0:40
0.2 Vf = 0:42
Vf = 0:44
0.1

0
0 1 2 3 4 5 6 7 8
Velocity [mm/s] #10!3

0.5
Vf = 0:33
0.4 Vf = 0:35
Area [mm2 ]

Vf = 0:37
0.3 Vf = 0:38
Vf = 0:40
0.2 Vf = 0:42
Vf = 0:44
0.1

0
0 0.05 0.1 0.15 0.2 0.25
Velocity / Vavg [-]

Figure 5.4: Velocity distribution for intra-tow domain

this value by the area covered by a single data point (An ), gives the total area for the
velocity of that histogram bin.

Av = An ∗ histcounts (un )

un
 (5.2)
Av = An ∗ histcounts
avg(u)

Figure 5.4 shows the velocity distribution within the tow. The intra-tow flow is very
uniform and slow (vintra < 1.0 µm/s). The normalized plot shows that the intra-tow
flow moves almost 100 times slower than the average flow.

The inter-tow velocity distribution is calculated using the same method. Line histograms
(where nbins = 50) are made of the velocity component u.
Chapter 5: Velocity Distributions 62

0.03 Vf = 0:33
Vf = 0:35
0.025
Vf = 0:37
Area [mm2 ]

0.02 Vf = 0:38
Vf = 0:40
0.015
Vf = 0:42
0.01 Vf = 0:44
0.005

0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4


Velocity [mm/s]

0.03 Vf = 0:33
0.025 Vf = 0:35
Area [mm2 ]

Vf = 0:37
0.02 Vf = 0:38
Vf = 0:40
0.015
Vf = 0:42
0.01 Vf = 0:44
0.005

2 4 6 8 10 12
Velocity / Vavg [-]

Figure 5.5: Velocity distribution for inter-tow domain

Figure 5.5 shows the inter-tow velocity distribution. Immediately, the effect of restricting
the flow by increasing the Vf is reflected in the absolute velocities. However, it is
interesting that the shape of the velocity distribution (shown by the normalized plot)
remains similar.

The velocity distributions of all the Vf show that within the intra-tow region, there is
a large area covered by slow flow, and then a small remainder of the area contributes
to the high-speed flow. The simulation reveals that even within the inter-tow domain,
there are multiple regimes of flow speeds. There is not simply the slow, intra-tow flow
and the fast inter-tow flow: the inter-tow flow is more complex.

This complexity arises due to the clearly defined flow channels that occupy only a small
region of the inter-tow domain. These consequences are presented in the following
section.
Chapter 5: Velocity Distributions 63

5.1.3 Volumetric Flow Rates

Section 5.1.1 showed that the intra-tow flow is slow and occupies a large percentage of
the total cross-sectional area. We are also interested in characterizing the mass transport
of the dual-scale flow. By doing this, we can understand how much of the resin flows
through which parts of the fabric.

The volumetric flow rate for a planar cross-section is defined as: Q = A ∗ v. We divide
the dual-scale flow into four flow rate regimes: the inter-tow regime, and the bottom,
middle, and top 33% of the velocity range within the inter-tow space.

Qhi = Q(u > 0.67 ∗ umax )


Qmed = Q(0.67 ∗ umax > u > 0.33 ∗ umax )
(5.3)
Qlow = Q(0.33 ∗ umax > u > uintra )
Qintra = Q(uintra )

Figure 5.6 contains two complimentary plots. In Figure 5.6a, the flow rate and area
occupied by each regime (normalized to unity) is shown. Figure 5.6b shows which
region of dual-scale fabric is occupied by each flow rate regime. The lighter the color,
the higher the volumetric flow rate.

Immediately we see that the intra-tow region occupies most of the flow area. However,
because it is orders of magnitude slower than the inter-tow flow, it contributes very little
(< 1%) to the total mass transport.

On the other hand, the inter-tow flow regimes, which occupy less than 30% of the total
area, contribute to more than 99% of the flow. A well defined flow-channel can be
established by considering the highest flow regime: Qhi . This regime occupies only 4%
of the total area, but almost 40% of the resin passes through it.

Figure 5.7 shows the same analysis performed on the plain weave simulations for the
entire range of fiber volume contents. The results are relatively uniform for all of the
fiber volume contents. Small deviations occur due to the discretization into the four
flow regions. The Vf = 0.33 & 0.35 flow rates have a higher (maximum) cutoff velocity,
resulting in their Qmed regime containing higher flow rates.

It is well known that in dual-scale fabrics resin flows much faster through the inter-tow
flow channels than in the porous tows [30]. However, these results give a quantitative
measure of this difference in flow between the intra- and inter-tow regions, as well as
within the inter-tow region itself.
Chapter 5: Velocity Distributions 64

Area

Intra-tow
Q cut > uintra
cut > 0.33
cut > 0.67

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


(a) Normalized area & volumetric flow rate

(b) Area occupied by different flow regimes

Figure 5.6: Comparison of area occupied vs. flow rates for the plain weave fabric
(Vf = 0.33)

Vf = 0:33
Qhi Qhi Vf = 0:35
Vf = 0:37
Vf = 0:38
Vf = 0:40
Qmed Qmed Vf = 0:42
Vf = 0:44

Vf = 0:33
Qlow Vf = 0:35 Qlow
Vf = 0:37
Vf = 0:38
Vf = 0:40
Qintra Vf = 0:42 Qintra
Vf = 0:44

0 0.1 0.2 0.3 0.4 0.5 0 0.2 0.4 0.6 0.8


Normalized volumetric .ow rate: Q / Qtot [-] Area ratio [-]

Figure 5.7: Flow rate regimes compared to area occupied for multiple Vf
Chapter 5: Comparison to Biaxial Fabric 65

Figure 5.8: A biaxial fabric generated with TexGen

5.2 Comparison to a Biaxial Fabric

In an effort to test the robustness of the dual-scale flow analysis framework developed
in this thesis, a second fabric has been analyzed. The same methodology, unit cell
generator, and simulation parameters that were applied for the plain weave fabric have
been used for a biaxial fabric. Because the fabric architecture and unit cell size are
significantly different, the absolute values of the velocities will differ for the two textiles.
Nevertheless, the shape of the velocity distributions can be compared.

The biaxial fabric consists of two layers of unidirectional, non-crimped fabric (NCF),
stacked on top of one another at 0°/90°. The material properties of the NCF are de-
scribed in Section 2.1.2, and the unit cells generated for the CFD simulation are listed
in Table 2.5. The simulation parameters are given in Table 4.1.

5.2.1 CFD Simulation

Unlike the plain weave fabric, the yarns of the biaxial fabric do not weave through
each other. The warp yarns run perfectly parallel to one another. This adds an extra
degree of symmetry to the unit cell, whose width can be cut in half and still represent
a repeating unit cell of the fabric. This symmetry is exploited in the flow model.

The biaxial mesh applies the same principles learned from the plain weave convergence
study: the inter-tow flow channels must have a high resolution in order to fully develop
the Navier-Stokes flow. The biaxial simulations all run with at least 1.5 million DOF.
However, they require only 1/3 the solution time compared to the plain weave simula-
tions. This is due to a faster convergence, which is a result of the biaxial flow field being
more uniform.
Chapter 5: Comparison to Biaxial Fabric 66

(a) Simulation mesh (b) Velocity field

Figure 5.9: Mesh and resulting flow field of biaxial fabric (Vf = 0.44)

5.2.2 Results

The architecture of the biaxial fabric is more uniform than that of the plain weave
(no weaving yarns). The flow channels of the biaxial fabric are unrestricted in passing
through the entire length of the unit cell, whereas the flow channels of the plain weave
must undulate through the woven yarns. This has a direct influence on the dual-scale
flow field, which is also more uniform. Almost the entire flow passes through the two
flow channels; the rest is blocked by a tightly packed intra-tow region.

Figure 5.10 shows the velocity distributions (Vf = 0.44 → 0.60) of the inter-tow flow.
The area y-axis is plotted on a logarithmic scale because of the large difference in area
covered by low and high velocities. The data becomes noisy for the small values of area,
likely because we are approaching the maximum resolution of the simulated velocity
field.

Compared to the plain weave, there is a much sharper decline in area covered by increas-
ing velocity. This means that the flow channel is more pronounced, or locally constrained
than in the plain weave fabric. In the plain weave, the inter-tow flow is more spread out
because the weaving yarns create more regions through which the flow can pass.

The volumetric flow rates in Figure 5.11 confirm the same trend. There is a stronger
discrepancy between the intra- and inter-tow regimes, and the drop from Qlow to Qhi is
also larger.

Comparing the velocity distributions of the biaxial and plain weave fabrics demonstrates
that the fabric architecture does have a strong influence on how the resin flows through
the fabric during an injection.
Chapter 5: Comparison to Biaxial Fabric 67

fvc44
10!1
fvc48
Area [mm2 ]

fvc50
fvc52
10!2 fvc54
fvc56
fvc58
fvc60
10!3
0.5 1 1.5 2 2.5 3
Velocity [mm/s]

fvc44
10!1 fvc48
Area [mm2 ]

fvc50
fvc52
fvc54
fvc56
10!2
fvc58
fvc60

5 10 15 20 25 30 35 40 45
Velocity / Vavg [-]

Figure 5.10: Biaxial fabric: velocity distributions for inter-tow domain

fvc44
Qhi Qhi fvc48
fvc50
fvc52
fvc54
Qmed Qmed fvc56
fvc58
fvc44 fvc60
fvc48
Qlow fvc50 Qlow
fvc52
fvc54
fvc56
Qintra fvc58 Qintra
fvc60

0 0.1 0.2 0.3 0.4 0.5 0 0.2 0.4 0.6 0.8 1


Normalized volumetric .ow rate: Q / Qtot [-] Area ratio [-]

Figure 5.11: Biaxial fabric: flow rate vs. area comparison for multiple Vf
Chapter 5: Comparison to Experiments 68

5.3 Comparison to Experiments

An experimental analysis of the velocity distributions in dual-scale fabrics has been


conducted at the CMAS Laboratory. My thanks goes Michael Deiss for making the
results as well as the data available in order to conduct this comparison between the
numerical and experimental velocity distributions. The full report can be found in [48].

5.3.1 Experimental Setup

In a CFD flow simulation, the velocity at each point in space is known. Getting such
localized data in reality is tricky, as an instrument that samples the flow might at the
same time obstruct it. The experiment uses particle tracking as a means of sampling
the velocity of the fluid through the dual-scale fabric.

The fluid is filled with particles, which can be detected as they pass through the fabric.
The developed RTM injection tool, shown in Figure 5.12, has round view holes on the
top and bottom plates. A microscope fitted with a high-speed camera is placed over the
view holes, which detects the particles as they pass through the field of view.

The high speed camera, a Photron UX 100, is set to record at 125fps. The frames are
first processed, then the difference between two frames is used to calculate how far (in
pixels) a particle has moved. The maximum resolution is 1 pixel, which means that a
particle moving too slow will not have traversed 1 pixel between frames. This results in
a minimum cut off speed of 0.272 mm/s.

Figure 5.12: RTM tool developed for dual-scale flow tracking [48]
Chapter 5: Comparison to Experiments 69

Table 5.1: Materials used in the particle tracking experiment

Textile Plain weave Aw = 280 g/m2


Fluid Ethylcinnamate µ = 0.0823 Pa s
Particles Sicastar black d = 10 µm

The materials used in the experiment (Table 5.1) were the basis for many of the pa-
rameters of the simulation performed in this thesis. The same plain weave fabric and
injection fluid used by the experiment were modelled in the simulation. This was chosen
purposefully so that a comparison between the two could be made.

5.3.2 Results

As in the simulated velocity distributions, the experiment also normalized the velocities
by an average velocity. The simulations normalize by the average velocity of the cross-
section. This value is not available in the experiment. Instead, it normalizes by the
“average porous flow velocity”, ṽav .

! !
1 wf low
ṽav = ṽP T V (5.4)
1 − Vf wf ab

The measurable quantity in the experiment is ṽP T V , which is the average velocity of
the fluid flowing where there are no fibers present. To find the average velocity in the
porous media, this value must be corrected by two terms. First, it is divided by the
porosity (p = 1 − Vf ) to account for the presence of the porous fabric. This increase to
ṽav is a result of no fluid being able to flow through the area occupied by the fibers. The
second correction comes from the experimental setup, which accounts for the difference
in widths between the empty cavity (wf low ) and where the fabric lies (wf ab ).

The particle tracking velocimetry (PTV) software DynamicStudio is used to analyze the
flow images and track the particles as they flow through the fabric. Histograms of the
resulting velocity distributions are shown in Figures 5.13 and 5.14.

There are a few caveats that need to be considered when comparing the simulation and
experimental results:

• Both the simulation and experiment give velocity histograms. However, the simu-
lation counts data points from the cross-sectional velocity field (ŷ-ẑ plane), whereas
the experiment is counting particles that are flowing through the length of the fab-
ric (x̂-ŷ plane).
Chapter 5: Comparison to Experiments 70

8000
Number of particles (n)
Experiment (Vf = 0:366)
6000

4000

2000

0
0 1 2 3 4 5 6 7 8 9 10
Velocity [mm/s]
Number of data points (un )

600
500 Simulation (Vf = 0:371)

400
300
200
100
0
0.05 0.1 0.15 0.2 0.25 0.3 0.35
Velocity [mm/s]

Figure 5.13: Comparison between absolute velocities (the markers represent the nu-
merical bins)

8000
Number of particles (n)

Experiment (Vf = 0:366)


6000

4000

2000

0
0 5 10 15 20 25
Normalized velocity: v=vav [-]
Number of data points (un )

600
500 Simulation (Vf = 0:371)

400
300
200
100
0
2 4 6 8 10 12
Normalized velocity: u=mean(utot ) [-]

Figure 5.14: Comparison between normalized velocity distributions (the markers


represent the numerical bins)
Chapter 5: Comparison to Experiments 71

• Also, the numerical data points come from a cross-section of the simulated steady-
state velocity field. The experimental data comes from particles tracked over a
period of time. In other words, the y-axis represents very different types of velocity
data points. This is the reason for the large difference in the number of data points.

• The binning of the data points also differs: the simulated data has nbins = 50
and the experimental data has nbins = 21. As different types of data are being
analyzed, forcing an equal number of bins is not sensible.

• Although the minimum cut off velocity of the image capturing setup is 0.272 mm/s,
the bias against low velocities extends into higher values. The experiment counts
few particles at v < 1.4 mm/s, even though regions of slow flow should exist. There
is uncertainty in the exact cause of the reduction in low-end velocities.

• A possible explanation comes from the fabric being fully saturated with a particle-
free Ethylcinnamate oil before the particle-filled oil is injected. This means that
the particles will primarily travel in the central regions of the flow channel (where
there is least resistance to flow). This would bias the particle sample towards
higher velocities. With additional time the oils will continue to mix, allowing for
particles to be tracked in the slow flowing regions as well.

In summary, this analysis is not intended to compare the specific values between the
experimental and simulated results, but rather give an insight to the shape or trend in
the velocity distribution. With this in mind, the analysis of both the numerical and
experimental histograms shows that a large number of points exhibit slow velocities,
and then the number decreases steadily into the regions of high velocity.
Chapter 5: Comparison with the State of the Art 72

5.4 Comparison with the State of the Art

5.4.1 Previous Models

Darcy’s Law is was originally derived for isotropic porous media and does not take into
account the dual-scale flow. Using Darcy’s Law, we can find the average velocity through
the entire cross-section of the porous medium. The flow rate is:

QDarcy = vdarcy ∗ Atot (5.5)

The Darcian flow rate has a linear dependence on the averaged velocity of the whole
porous medium. As the cross-sectional area increases, so does the volumetric flow rate.
However, this is not accurate for a dual-scale textile. By averaging the velocity, the
difference between the fast inter-tow flow and the slow intra-tow flow is not considered.

Section 1.3.1 introduced the current dual-scale flow models which make use of the vol-
ume averaging technique. When considering dual-scale textiles, the volume averaging is
carried out on two different length-scales: the tow-level and fiber-level [24]. By splitting
the flow into the inter- and intra-tow regions, both regimes are considered independently.
Nevertheless, the volumetric flow rate still has a linear dependence on the volume aver-
aged velocities.

QSoA = hvinter i ∗ Ainter + hvintra i ∗ Aintra (5.6)

The state of the art method is a significant improvement over Darcy’s Law, especially
when dual-scale fabrics are being considered. However, the numerical efficiency of this
method comes at the cost of a loss in detailed information about the local flow variations,
which are significant in the inter-tow region.

5.4.2 Dual-Scale Flow Simulation

The previous results in this chapter clearly show that the volumetric flow rate in a
dual-scale fabric increases strongly in the region of the flow channels. The form of this
increase can be analyzed by plotting the flow rate as a function of area sampled around
the flow channel. The benefit of performing the Navier-Stokes simulation for the dual-
scale flow becomes clear: we can analyze the localized velocity changes occurring in the
inter-tow flow channels.
Chapter 5: Comparison with the State of the Art 73

Figure 5.15: Visualization of the area selection around flow channels. The contour
lines show the top 20%, 40%, 60%, and 80% velocity regions. The (zagged) black line
marks the intra-inter-tow boundary.

1.2
Normalized volumetric .ow rate [-]

0.8

0.6

0.4
Plain Weave Simulation
Darcy's Law
0.2 State of the Art
Intra-inter boundary
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Fraction of total cross-sectional area [-]

Figure 5.16: Comparison between dual-scale flow rate models (velocity data from
Vf = 0.33 simulation)

The contour lines in Figure 5.15 mark the top 20%, 40%, 60%, and 80% velocity regions
(considering both intra- and inter-tow domains). In Figure 5.15 the velocity has been
discretized into five regions, whereas in Figure 5.16 it is split into 200. As the number
of discretizations approaches infinity, the functional form of the volumetric flow rate is
achieved.

The volumetric flow rate in Figure 5.16 is calculated by scanning an increasing fraction of
the total cross-sectional area, beginning with the highest velocity region. The volumetric
flow rate for each local velocity data point is calculated, and summed:

n
X m
X
QSimulation = v i ∗ Ai + v j ∗ Aj (5.7)
Ainter Aintra

The analysis in Figure 5.16 shows a clear non-linear behavior of the volumetric flow
rate. The initial increase in the mass transport is very steep at the central regions of the
flow channels. Already at the first contour line (top 20%), almost all the mass transport
Chapter 5: Comparison with the State of the Art 74

is accounted for. And once the intra-inter-tow boundary is crossed, there is almost no
additional increase in the volumetric flow rate. This analysis shows the same physical
effect as is seen in Figures 5.6 and 5.7.

1.2
Normalized volumetric .ow rate [-]

0.8

0.6

0.4
Biaxial Simulation
Darcy's Law
0.2 State of the Art
Intra-inter boundary
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Fraction of total cross-sectional area [-]

Figure 5.17: Comparison between dual-scale flow rate models and Biaxial simulation
(velocity data from Vf = 0.50 simulation)

Figure 5.17 presents the same analysis for the velocity data of the biaxial simulation.
The result shows a similar non-linear increase in the volumetric flow rate, which deviates
from the linear models.

1.2
Normalized volumetric .ow rate [-]

0.8

0.6

0.4

Plain weave simulation (Vf = 0:40)


0.2 Biaxial simulation (Vf = 0:44)
State of the Art
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Fraction of inter-tow cross-sectional area [-]

Figure 5.18: Comparison between the plain weave and biaxial simulations for the
inter-tow flow rates

Figure 5.18 compares the plain weave and biaxial volumetric flow rates for the inter-tow
region. The slope of the increase of the biaxial data is steeper than that of the plain
weave. This is a direct result of the biaxial flow channels being more localized than
Chapter 5: Comparison with the State of the Art 75

those of the plain weave. The increased localization is quantified in Figures 5.10 and
5.11.

The remarkable result of this analysis is that the mass transport of the fluid through
the complex structure of a dual-scale fabric has been reduced to such a smooth curve.
The dual-scale flow of different types of fabrics can be modelled and compared with
the volumetric flow rate analysis. This lends itself well for becoming the basis of a new
mathematical formulation for describing the dual-scale flow through composite fabrics.
These possibilities will be explored further in Chapter 6.
Chapter 5: Equivalent Flow Channel Model 76

5.5 Equivalent Flow Channel Model

Section 5.4 introduced the idea that the Navier-Stokes simulation offers insightful infor-
mation about the volumetric flow rates in a dual-scale fabric. This has a high potential
of being used to make a mathematical model which incorporates the effects of the dual-
scale saturated flow.

The idea is to transform the complex inter-tow flow into an “equivalent flow channel”
using the volumetric flow rates. Conceptually, this is similar to the volume averaging
techniques, which homogenizes the inter-tow flow of an entire meso-scale unit cell. How-
ever, with the equivalent flow channel model, the level of detail sampled by the inter-tow
channel can be increased up to the maximum resolution of the Navier-Stokes simulation.

We begin by considering the four discretized flow rate regimes presented in Section 5.1.3:
Qintra , Qlow , Qmed , and Qhi . These four regimes contain the total volumetric flow rate
of the entire cross section. The equivalent flow channel is used to model the inter-tow
flow only.

Using the principle of mass conservation for an incompressible fluid in a closed system
(i.e. saturated flow), we require that the mass transport of the equivalent flow channel
Qeq.ch. is equal to that of the inter-tow region.

Qeq.ch. ≡ Qinter
(5.8)
Qeq.ch. = Qlow + Qmed + Qhi

The volumetric flow rate can be expressed as the surface integral of the velocity passing
through the area under consideration. When the flow is normal to a planar cross-section,
this reduces to:
ZZ
Q= v · dA → Q=v∗A (5.9)
A

Applying the simplified definition to the equivalent channel flow rate results in:

veq.ch. ∗ Aeq.ch. = Qlow + Qmed + Qhi (5.10)

Rewriting the inter-tow flow rates as a sum and solving (5.10) for the equivalent velocity
gives:
3
X
Qi
i
veq.ch. = (5.11)
Aeq.ch.
Chapter 5: Equivalent Flow Channel Model 77

which states that the velocity of fluid in the equivalent flow channel veq.ch. is equal to
the volumetric flow rate divided by the available flow area of the channel Aeq.ch. .

Now we would like to apply the concept of the equivalent flow channel to a model of
the dual-scale flow: one that considers both the intra- and inter-tow regimes. Beginning
again with the conservation of fluid mass for an incompressible fluid in a closed system
(i.e. saturated flow), we write:

Qtot = Qintra + Qinter (5.12)

Considering a planar cross-section Atot of the unit cell, we can equate the average fluid
velocity vf luid to the total volumetric flow rate, divided by the available flow area.

3
X
Qintra + Qi
i
vf luid = (5.13)
Atot (1 − Vf )

The available flow area is the area not occupied by fibers, which is given by the porosity
(p = 1 − Vf ) of the dual-scale fabric.

Discretizing the inter-tow flow rate into three regions is an arbitrary choice. If we
consider the state of the art comparison plot (Figure 5.16), we see that rather than
just being split into three regions, the inter-tow volumetric flow rate has a functional
dependence on the area. Re-expressing (5.13) with the functional dependence gives:

Qintra + f (Qinter , Ainter )


vf luid = (5.14)
Atot (1 − Vf )

This formulation is advantageous because it provides a single value for the saturated
flow velocity, while still considering the highly non-linear volumetric flow rate of the
dual-scale fabric. It also bridges the functional dependence of the volumetric flow rate
to the average fluid velocity, which is used in Darcy’s Law.
Chapter 6

Conclusions and Outlook

Within the scope of this thesis, a framework for analyzing the dual-scale flow in compos-
ite textiles has been developed. This framework includes: the unit cell generator which
uses TexGen to build 3D periodic unit cells, the COMSOL CFD analysis which simu-
lates the saturated dual-scale flow through the generated unit cells, and the MATLAB
analysis which characterizes the velocity distributions and volumetric flow rates.

The resulting analyses of the dual-scale flow show that large variations in the flow ve-
locity exist. By looking beyond the well-established difference between intra- and inter-
tow flow, the dual-scale simulations also show significant variations occurring within the
inter-tow flow. These local variations have an influence on both the velocity distribution
and the mass transport of the fluid in the dual-scale textile.

The comparison to the state of the art models demonstrates that a consideration of
the inter-tow flow variations can be used to provide a more accurate description of the
dual-scale flow. It has been shown that the textile architecture has a strong influence
on the increase in the inter-tow volumetric flow rate.

The equivalent flow channel model offers a mathematical description of the saturated
flow velocity which includes the local variations of the inter-tow flow. Due to the well-
behaved nature of the volumetric flow rate curve, this analysis has the potential to form
the basis of a new analytical model for dual-scale saturated flows.

78
Chapter 6: Conclusions and Outlook 79

6.1 Further Work

As with most endeavours, additional improvements can always be made.

The dual-scale flow simulation models the saturated, steady-state flow in the textile. A
logical next step would be to model the unsaturated, transient flow. However, this adds
complexity to the numerical model. Not only are the time-dependent Navier-Stokes
equations more computationally expensive to solve, but the two-phase unsaturated flow
requires additional considerations (impregnation of tows, formation of pores). To com-
pensate for the added complexity, a reduction in the model resolution is likely necessary.

The compaction of the fabric has been implemented in a purely geometrical fashion.
The nesting of the unit cell and reduction of the tow height emulates the compression
of the fabric. There is no mechanical consideration of the compaction. Adding this
would be a very nice improvement to the textile model: first, to simulate the viscoelastic
compaction [69] of the textile unit cell due to an applied pressure, and then run the CFD
simulation with the compacted unit cell. A mechanical compaction of the fabric would
add a higher degree of realism to the textile model without influencing the complexity
of the flow simulation.

The equivalent flow channel model has been developed from the discussions and ideas
resulting from this numerically focused work. For a proper analytical treatment, it
should be verified with simulations and experimental results.

6.2 Potential Applications

The highly non-linear increase of the volumetric flow rate in the inter-tow region also
offers valuable insights. Knowing that 70% of the fluid passes through only 10% of the
cross-sectional area (for the plain weave fabric) can be used to fine tune the processing
parameters in LCM applications.

For particle-filled resin injections, knowing where and in what concentration the particles
will be deposited is beneficial. The injection parameters can be optimized to distribute
particles to the intra- and inter-tow space as required, which results in reduced particle
filtration effects and an efficient increase in the composite’s matrix properties.

The highly localized flow channels could also be exploited to create new kinds of self-
healing composites. Instead of embedding hollow fibers in the laminate [42], the flow
channels could be made hollow by the injection of a second, solvable fluid. This way,
Chapter 6: Conclusions and Outlook 80

the vascular network is integrated directly into the flow channels of the matrix, without
having to add any additional components.

Finally, the improved mathematical model of the dual-scale saturated flow could be
implemented into an LCM filling software for entire (macro-scale) composite parts. A
more refined model of the saturated flow would improve the accuracy of the entire filling
simulation, as the progression of the unsaturated flow front depends on the supply of
resin from the saturated flow. Increasingly accurate filling models will allow for a higher
degree of optimization of industrial-grade composite parts.
Bibliography

[1] Paolo Ermanni. Composites technologien. ETH University Lecture Script, 2007.

[2] Robert M Jones. Mechanics of composite materials. CRC Press, 1998.

[3] Roger Bacon and Charles T Moses. Carbon Fibers, from Light Bulbs to Outer
Space. In High Performance Polymers: Their Origin and Development, pages 341–
353. Springer, 1986.

[4] The History of Carbon Fiber. URL http://www.hj3.com/company/


history-of-carbon-fiber/.

[5] Giovanni Terrasi. Engineering design with polymers and advanced composites.
ETH University Lecture Script, 2014.

[6] Hasan Salek and Paul Trudeau. Development of low cost fuselage frames by resin
transfer molding, 09 2013. URL http://dx.doi.org/10.4271/2013-01-2325.

[7] Branko Sarh Billy Roeseler and Max Kismarton. Composite Structures – The First
100 Years. Kyoto, Japan, 2007. ICCM-16.

[8] Boeing. AERO Magazine QTR04 06, 2006.

[9] Robert H. Holley. The Great Metal Tube in the Sky. URL http://arch5541.
wordpress.com/2013/01/08/the-great-metal-tube-in-the-sky/.

[10] Durrer Martin. Feasibility study for a composite aircraft fuselage section made by
an integral design. volume 6th Technical Conference of Swiss SAMPE Technical
Conference. SAMPE, 2013.

[11] Babak Ganjeh and Mohd Roshdi Hassan. Cost-efficient composite processing tech-
niques for aerospace applications–a review. Applied Mechanics and Materials, 325:
1465–1470, 2013.

[12] David Fraser Hank Kleespies Cliff Vasicek Cyril Annarella. Spacecraft Structures.
Purdue University Lecture Script, 1991.

81
Bibliography 82

[13] Abdel A Abusafieh, Dan R Federico, Steven J Connell, Eri J Cohen, and Paul B
Willis. Dimensional stability of cfrp composites for space-based reflectors. In Inter-
national Symposium on Optical Science and Technology, pages 9–16. International
Society for Optics and Photonics, 2001.

[14] Mario Danzi. Out of autoclave processing. ETH University Lecture Script, 2012.

[15] Ginger Gardiner. Out-of-autoclave prepregs: Hype or revolution,


January 2011. URL http://www.compositesworld.com/articles/
out-of-autoclave-prepregs-hype-or-revolution.

[16] M Deleglise, C Binetruy, and P Krawczak. Simulation of LCM processes involving


induced or forced deformations. Composites Part A: applied science and manufac-
turing, 37(6):874–880, 2006.

[17] Christian Fais. Lightweight automotive design with hp-rtm. Reinforced Plastics,
55(5):29–31, 2011.

[18] JA Molnar, L Trevino, and LJ Lee. Liquid flow in molds with prelocated fiber mats.
Polymer Composites, 10(6):414–423, 1989.

[19] WD Brouwer, ECFC Van Herpt, and M Labordus. Vacuum injection moulding for
large structural applications. Composites Part A: Applied Science and Manufac-
turing, 34(6):551–558, 2003.

[20] Dominik Bender, Jens Schuster, and Dirk Heider. Flow rate control during vacuum-
assisted resin transfer molding (vartm) processing. Composites Science and Tech-
nology, 66(13):2265–2271, 2006.

[21] Y Wang and SM Grove. Modelling microscopic flow in woven fabric reinforcements
and its application in dual-scale resin infusion modelling. Composites Part A:
Applied Science and Manufacturing, 39(5):843–855, 2008.

[22] Chung Hae Park, Aurélie Lebel, Abdelghani Saouab, Joël Bréard, and Woo Il
Lee. Modeling and simulation of voids and saturation in liquid composite molding
processes. Composites Part A: Applied Science and Manufacturing, 42(6):658–668,
2011.

[23] CC Wong, AC Long, M Sherburn, F Robitaille, P Harrison, and CD Rudd. Com-


parisons of novel and efficient approaches for permeability prediction based on the
fabric architecture. Composites Part A: Applied Science and Manufacturing, 37(6):
847–857, 2006.
Bibliography 83

[24] Hua Tan and Krishna M Pillai. Multiscale modeling of unsaturated flow in dual-
scale fiber preforms of liquid composite molding I: Isothermal flows. Composites
Part A: Applied Science and Manufacturing, 43(1):1–13, 2012.

[25] Bart Verleye, Stepan Vladimirovitch Lomov, Andrew Long, Ignace Verpoest, and
Dirk Roose. Permeability prediction for the meso–macro coupling in the simulation
of the impregnation stage of resin transfer moulding. Composites Part A: Applied
Science and Manufacturing, 41(1):29–35, 2010.

[26] Krishna M Pillai. Governing equations for unsaturated flow through woven fiber
mats. part 1. isothermal flows. Composites Part A: Applied Science and Manufac-
turing, 33(7):1007–1019, 2002.

[27] Romain Brault, Michel Niquet, and Sébastien Mistou. Experimental and numerical
modelling of LRI process. Key Engineering Materials, 446:121–130, 2010.

[28] ND Ngo and KK Tamma. Microscale permeability predictions of porous fibrous


media. International Journal of Heat and Mass Transfer, 44(16):3135–3145, 2001.

[29] Bart Verleye. Computation of the permeability of multi-scale porous media with ap-
plication to technical textiles. Katholieke Universiteit Leuven, PhD Thesis, ISBN,
pages 978–90, 2008.

[30] Krishna M Pillai. Modeling the unsaturated flow in liquid composite molding pro-
cesses: a review and some thoughts. Journal of Composite materials, 38(23):2097–
2118, 2004.

[31] November 2014. URL http://www.ifam.rwth-aachen.de/aw/cms/IFAM/Themen/


forschung/~vps/mehrskalige-modellierung-des-nichtlinear/?lang=en.

[32] Richard S Parnas and FR Phelan. The effect of heterogeneous porous media on
mold filling in resin transfer molding. Sampe Quarterly, 22(2):53–60, 1991.

[33] Baiju Z Babu and Krishna M Pillai. Experimental investigation of the effect of fiber-
mat architecture on the unsaturated flow in liquid composite molding. Journal of
Composite Materials, 38(1):57–79, 2004.

[34] BR Gebart. Permeability of unidirectional reinforcements for RTM. Journal of


Composite Materials, 26(8):1100–1133, 1992.

[35] Chih-Yuan Chang and Lih-Wu Hourng. Numerical simulation for the transverse im-
pregnation in resin transfer molding. Journal of reinforced plastics and composites,
17(2):165–182, 1998.
Bibliography 84

[36] Ph Vandeurzen, Jan Ivens, and Ignace Verpoest. A three-dimensional micromechan-


ical analysis of woven-fabric composites: I. geometric analysis. geometric analysis.
Composites Science and Technology, 56(11):1303–1315, 1996.

[37] Martin Sherburn. Geometric and mechanical modelling of textiles. PhD thesis,
University of Nottingham, 2007.

[38] Jonathan Josiah Crookston. Prediction of elastic behaviour and initial failure of
textile composites. PhD thesis, University of Nottingham, 2004.

[39] Baoxing Chen and Tsu-Wei Chou. Compaction of woven-fabric preforms in liquid
composite molding processes: single-layer deformation. Composites Science and
Technology, 59(10):1519–1526, 1999.

[40] Markus Nordlund, SP Fernberg, and TS Lundström. Particle deposition mech-


anisms during processing of advanced composite materials. Composites Part A:
Applied Science and Manufacturing, 38(10):2182–2193, 2007.

[41] Elisabete F Reia da Costa, Alexandros A Skordos, Ivana K Partridge, and Amir
Rezai. Rtm processing and electrical performance of carbon nanotube modified
epoxy/fibre composites. Composites Part A: Applied Science and Manufacturing,
43(4):593–602, 2012.

[42] BJ Blaiszik, SLB Kramer, SC Olugebefola, Jeffrey S Moore, Nancy R Sottos, and
Scott R White. Self-healing polymers and composites. Annual Review of Materials
Research, 40:179–211, 2010.

[43] Michael Deiss. Characterization of pore-scale effect in textiles during LCM pro-
cessing. Technical report, ETH Zürich, March 2014.

[44] Louise P. Brown. Texgen open source textile generator. URL http://texgen.
sourceforge.net/index.php/Main_Page.

[45] Hua Lin, Louise P Brown, and Andrew C Long. Modelling and simulating textile
structures using texgen. Advanced Materials Research, 331:44–47, 2011.

[46] Mehmet Karahan, Stepan V Lomov, Alexander E Bogdanovich, Dmitri Mungalov,


and Ignaas Verpoest. Internal geometry evaluation of non-crimp 3d orthogonal
woven carbon fabric composite. Composites Part A: Applied Science and Manufac-
turing, 41(9):1301–1311, 2010.

[47] P Potluri and TV Sagar. Compaction modelling of textile preforms for composite
structures. Composite Structures, 86(1):177–185, 2008.
Bibliography 85

[48] Michael Deiss. Experimental investigation of the velocity distribution of saturated


resin flow during LCM. Master’s thesis, ETH Zürich, March 2015.

[49] Suter-Kunstoffe AG. Composite materials product catalogue, December 2014.

[50] Hana Zrida, Erik Marklund, Zoubir Ayadi, and Janis Varna. Effective stiffness of
curved 0 ◦ -layers for stiffness determination of cross-ply non-crimp fabric compos-
ites. Journal of reinforced plastics and composites, 33(14):1339–1352, 2014.

[51] SAERTEX GmbH & Co. U-C-PB-336g/m2 data sheet.

[52] F Robitaille, AC Long, IA Jones, and CD Rudd. Automatically generated geometric


descriptions of textile and composite unit cells. Composites Part A: Applied Science
and Manufacturing, 34(4):303–312, 2003.

[53] Oxford Dictionary of English. Oxford University Press, 2012. Porous.

[54] Henry Darcy, Henry Darcy, and Henry Darcy. Les fontaines publiques de la ville de
Dijon. 1856.

[55] Craig T Simmons. Henry Darcy (1803–1858): Immortalised by his scientific legacy.
Hydrogeology Journal, 16(6):1023–1038, 2008.

[56] Frank M White. Fluid Mechanics (2011). McGraw-Hill, Seventh edition, 2003.

[57] MT Senoguz, FD Dungan, AM Sastry, and JT Klamo. Simulations and experi-


ments on low-pressure permeation of fabrics: Part II—the variable gap model and
prediction of permeability. Journal of composite materials, 35(14):1285–1322, 2001.

[58] PC Carman. Fluid flow through granular beds. Transactions-Institution of Chem-


ical Engineeres, 15:150–166, 1937.

[59] Alexander L Berdichevsky and Zhong Cai. Preform permeability predictions by


self-consistent method and finite element simulation. Polymer Composites, 14(2):
132–143, 1993.

[60] Z Cai and AL Berdichevsky. An improved self-consistent method for estimating


the permeability of a fiber assembly. Polymer composites, 14(4):314–323, 1993.

[61] Josias Van der Westhuizen and J Prieur Du Plessis. An attempt to quantify fibre
bed permeability utilizing the phase average navier-stokes equation. Composites
Part A: Applied Science and Manufacturing, 27(4):263–269, 1996.

[62] COMSOL Multiphysics. CFD Module User’s Guide, COMSOL 4.4 edition, Novem-
ber 2013.
Bibliography 86

[63] Francisco J Valdes-Parada, J Alberto Ochoa-Tapia, and Jose Alvarez-Ramirez. On


the effective viscosity for the darcy–brinkman equation. Physica A: Statistical Me-
chanics and its Applications, 385(1):69–79, 2007.

[64] COMSOL Multiphysics. COMSOL Multiphysics Reference Manual, COMSOL 4.4


edition, November 2013.

[65] COMSOL Multiphysics. What does degrees of freedom (DOFs) mean in COMSOL
Multiphysics. URL http://www.comsol.com/support/knowledgebase/875/.

[66] David Wells. Are these the most beautiful? The Mathematical Intelligencer, 12(3):
37–41, 1990.

[67] ETH IT Services. Introducing EULER, . URL http://www.clusterwiki.ethz.


ch/brutus/Introducing_EULER.

[68] ETH IT Services. Comsol Multiphysics, . URL http://www.clusterwiki.ethz.


ch/brutus/Comsol_Multiphysics.

[69] PA Kelly, R Umer, and S Bickerton. Viscoelastic response of dry and wet fibrous
materials during infusion processes. Composites Part A: Applied Science and Man-
ufacturing, 37(6):868–873, 2006.

[70] Salvatore P Sutera and Richard Skalak. The history of Poiseuille’s law. Annual
Review of Fluid Mechanics, 25(1):1–20, 1993.

[71] Flow Between Parallel Plates. URL http://farside.ph.utexas.edu/teaching/


336L/Fluidhtml/node108.html.

[72] Grigorios Koutsoukis. Flow Monitoring and Permeability Mapping in Composite


Processing. ETH Zürich, 2013.
Appendix A

Official Documents

This appendix contains the official documents which form the contractual agreement of
this master’s thesis. The documents are listed in the following order:

• ETH Zürich Declaration of Originality (signed)

• CMAS Task Assignment (signed)

87
Eidgenössische Technische Hochschule Zürich Department of Mechanical and Process Engineering
Swiss Federal Institute of Technology Zurich Laboratory of Composite Materials and Adaptive Structures

ETH Zurich
Leonhardstrasse 21
CH-8092 Zurich

Stephan Krüsi

Student: ETH-Nr: 12-945-333 Departement: MAVT

Hochschule (if external student):


Title: Numerical analysis and modelling of the velocity distribution of
Thesis: saturated resin flow during LCM
Kind of Thesis: MA Semester: HS 2014

Supervisor: Prof. Dr. P. Ermanni

Advisors: Jesus Maldonado

Start of the work: 06/10/2014

Intermediate presentation (Zwischenpräsentation): 12/12/2014

Final presentation (Endpräsentation): 20/03/2015

Deadline delivery final report: 10/04/2015

Introduction
Liquid Composite Molding (LCM) is a technique for manufacturing fiber-reinforced composite
materials. In LCM, a dry preform is placed into a cavity and a liquid polymer, such as a
thermoset resin, is injected under pressure or infused under vacuum. The part is demolded
after impregnation of the preform and cure of the resin. Darcy’s Law is used as a basis for
modelling the macroscopic flow in engineering textiles: it defines a pressure-driven, volume-
averaged flux of a newtonian fluid through a permeable porous medium. However, Darcy’s
law is a simplification that assumes a homogeneous velocity distribution inside the flow
channels of the porous medium. Homogeneous, in this context, means that the resin moves
at a constant velocity along the cross-section of the textile during impregnation.

This thesis will contribute to a research project which is, in general terms, concerned with the
flow of resins through textiles for production of fiber-reinforced composites. The research
project focuses on revising the overall assumption of a homogeneous velocity distribution
inside the flow channels. These channels are the pores in the textile medium for resin flow,
and when considering this level of magnification, the term “pore-scale” is henceforward used.
In summary, the literature research in composite processing shows that pore-scale modelling
becomes an increasingly important topic. Pore-scale effects have already been shown to
exist in tubular flow channels [1-3], and consequences of them have been experimentally
noticed [4, 5]. The dual scale properties of textiles are also a consideration, due to the known
fact that initially injected mostly remains inside the fiber bundles. Tan and Pillai presented a
methodology coupling macroscopic flow simulations with microscopic simulations of fiber
bundle impregnation to account for dual-scale properties of textiles [6]. Studies that focus on
the pore-scale geometry in textiles have already been conducted, but their aim remains to
derive the flux within a unit cell to obtain a representative textile permeability, as presented

IDMF-CMAS/Kruesi - MA - Aufgabenstellung.docx Page 1


by e.g. Verleye et al [7] and Mahadik et al [8]. However, to apply such unit-cell numerical
models into industrial-scale modelling for composite production would be an extremely time-
intensive and resource-consuming task, rendering them impractical. The aim in this research
project is the development of a macroscopic model which considers the effects of pore-scale
flow in textiles within reasonable model complexity.

This thesis will provide a numerical modelling component to support the overall aim of the
research project. From the experimental side in the research project, previous work at the
CMAS Lab (Figure 1) shows that resin moves at significantly different velocities within the
textile, and this depends strongly on the textile architecture. Figure 1 shows an RTM
experiment where a red-colored resin is injected right after a transparent resin: the image on
the left is a photo taken right before the red resin enters the textile, and the right image
shows the experiment after 7 minutes. It can be seen how the overall flow front has
advanced a small distance compared to the distance travelled by the red resin, showing the
large difference in velocity. From the numerical side, some models on a macroscopic scale
have been derived at CMAS, but their definition of resin velocity is not currently linked to the
textile geometry, thus the need to numerically determine this dependency. The thesis work
will focus on simulating the flow through unit cells, emulating the situation in a textile. The
aim is to characterize the variability in resin movement velocity according to the textile
architecture. Factors such as fiber volume content, layup, and bundle size are to be
investigated.

Figure 1: Experiments performed at the CMAS Lab to investigate the saturated flow of resins
during RTM, showing the mixture of resin colors due to velocity variations, which are
currently not considered in state-of-the-art models and simulations.

Objectives
In this Master’s thesis, the aim is to develop a numerical simulation model for investigating
flow during composite processing and its dependency on textile architecture. This model will
serve as a tool to understand better the flow within a representative unit cell. An automated
geometry construction in the simulation should also be considered to analyze the effect of
different textile architectures on the velocity profile of the resin.

Work breakdown
In order to address these objectives, following work-packages have to be carried out:

1) Literature research about pore-scale flow and flow modelling in textiles.


2) Construction of a 2D numerical simulation, based on Darcy’s law and other models
from literature, as an initial reference for flow in textiles.
3) Development of a 3D numerical flow simulation with a representative unit-cell for flow
in textiles.

IDMF-CMAS/Kruesi - MA - Aufgabenstellung.docx Page 2


4) Addition of an automated geometry construction and post-processing system into the
3D flow simulation, in order to determine the dependency of resin flow on the textile
architecture.

The organization of the work is depicted in following table. The deadlines are also shown.

Week 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27

02. Mrz
09. Mrz
16. Mrz
23. Mrz
30. Mrz
06. Apr
03. Nov
10. Nov
17. Nov
24. Nov
01. Dez
08. Dez
15. Dez
22. Dez
29. Dez
05. Jan
12. Jan
19. Jan
26. Jan
02. Feb
09. Feb
16. Feb
23. Feb
06. Okt
13. Okt
20. Okt
27. Okt
Date
Tasks
Literature research
2D modelling
Intermediate presentation
3D unit cell modelling
3D automated modelling
Thesis writing
Final presentation

Bibliography

[1] Louis B, Maldonado J, Klunker F, Ermanni P. Characterization of Resin Flow in a Flow


Capillary During High Temperature LCM Injection. Society for the Advancement of Material
and Process Engineering, Baltimore, MD, USAMay 21-24, 2012.

[2] Louis B, Maldonado J, Klunker F, Lämmlein T, Ermanni P. Characterization of Reactive


Resin Flow During High Temperature LCM Injection. SAMPE EUROPE 33rd International
Technical Conference (SEICO 2012), Paris, FranceMarch 26-27, 2012.

[3] Maldonado J, Louis B, Klunker F, Ermanni P. Reactive flow of thermosetting resins -


Implications to LCM processing. 11th International Conference Flow Processing in
Composite Materials (FPCM-11), Auckland, New Zealand, University of Auckland, Faculty of
EngineeringJuly 9-12, 2012.

[4] Maldonado J, Müller S, Klunker F, Ermanni P. Curing inhomogeneity during reactive flow
in fiber preforms. SAMPE US, Long Beach, California2013.

[5] Owens JN. Variable catalyst injection rate for cycle time reduction in liquid composite
molding. The 9th International Conference on Flow Processes in Composite Materials,
Montréal (Québec), Canada2008.

[6] Tan H, Pillai KM. Multiscale modeling of unsaturated flow in dual-scale fiber preforms of
liquid composite molding III: reactive flows. Composites: Part A. 2012;43:29-44.

[7] Verleye B, Lomov SV, Long A, Verpoest I, Roose D. Permeability prediction for the
meso–macro coupling in the simulation of the impregnation stage of Resin Transfer
Moulding. Composites: Part A. 2010;41:29-35.

[8] Mahadik Y, Hallett SR. Finite element modelling of tow geometry in 3D woven fabrics.
Composites: Part A. 2010;41:1192-1200.

IDMF-CMAS/Kruesi - MA - Aufgabenstellung.docx Page 3


Appendix B

Unit Cell Generator Code

The full source code of UnitCellGenerator.py is given in this appendix. It was


developed using TexGen 3.7.0 and is fully compatible with the newest version, TexGen
3.8.0.

A detailed explanation of the unit cell generator is given in Section 2.2.2 and the code
diagram can be found in Figure 2.10.

93
 
 

#################################################################################
# This TexGen code has been written as part of an ETH Zurich Master's thesis. #
# If used, please cite: #
# Stephan Krusi. Numerical Analysis and Modelling of the Velocity Distribution #
# of Saturated Resin Flow during LCM. Master's thesis, ETH Zurich, 2015. #
# #
# This program is distributed in the hope that it will be useful, #
# but WITHOUT ANY WARRANTY; without even the implied warranty of #
# MERCHANTABILITY or FITNESS FOR A PARTICULAR PURPOSE. #
# #
# This code generates periodic unit cells of dual-scale textiles for CFD #
# modelling. A plain weave, biaxial, and twill fabric have been implemented. #
#################################################################################

#####################
# FABRIC TYPE #
#####################

# USER INPUT: Choose the type of fabric architecture:


fabricType = "PW"
# fabricType = "BIAXIAL"
# fabricType = "TWILL"

print("Generating a "+ str(fabricType) +" fabric.")

# Export Switches
exportSwitchSTEP = True
exportSwitchMESH = False
exportSwitchPNG = False

#####################
# MATERIAL DATA #
#####################

if fabricType == "PW":

# Fabric
arealWeight = 280.0 # [g/m^2]
fiberDensity = 2.58 # [g/cm^3] E-glass

# Warp
warpTexStr = "EC9-68x3 t0"
warpTexNum = 204.0 # [g/1000m]
warpFibDiameter = 9.0 # [um] diameter of 1 fiber filament
warpCount = 7.0 # [warps/cm]
warpSpacing = 10./warpCount # [mm] spacing between two warps in [mm]
warpWidth = 1.1855 # [mm] warp width: from IMAGE ANALYSIS data
warpGap = warpSpacing - warpWidth # [mm] space in 1 warp period not occupied

# Weft
weftTexStr = "EC9-204"
weftTexNum = 204.0 # [g/1000m]
weftFibDiameter = 9.0 # [um]
weftCount = 6.5 # [warps/cm]
weftSpacing = 10./weftCount # [mm] spacing between two wefts in [mm]
weftWidth = 1.3469 # [mm] weft width: from IMAGE ANALYSIS data
weftGap = weftSpacing - weftWidth # [mm] space in 1 weft period not occupied

# Cross-section exponent for Power ellipse


crosssectionPower = 1.8 # from 2007 Sherburn

# Give an initial height


fabThickness = 0.385 # FVC_tows_avg = 0.51214, FVC_textile = 0.40094
zshift = -0.03
textileResolution = 24

# Parameters for unit cell

    1
UnitCellGenerator.py  
   
 
 

numWarps = 2 # number of warp yarns


numWefts = 2 # number of weft yarns

# For layered textiles / nesting


numLayers = 3

elif fabricType == "BIAXIAL":

# Fabric
arealWeight = 313 # [g/m^2]
fiberDensity = 1.60 # [g/cm^3] CARBON

# Warp
warpTexStr = "TORAY T620SC 50C 24K"
warpTexNum = 1370.2 # [g/1000m]
warpFibDiameter = 6.7403 # [um] diameter of 1 fiber filament
warpCount = 2.1947 # [warps/cm]
warpSpacing = 10./warpCount # [mm] spacing between two warps in [mm]
warpWidth = 4.2780 # [mm] warp width: from IMAGE ANALYSIS data
warpGap = warpSpacing - warpWidth # [mm] space in 1 warp period not
occupied

# Weft
weftTexStr = "TORAY T620SC 50C 24K"
weftTexNum = 1370.2 # [g/1000m]
weftFibDiameter = 6.7403 # [um] diameter of 1 fiber filament
weftCount = 2.1947 # [warps/cm]
weftSpacing = 10./warpCount # [mm] spacing between two warps in [mm]
weftWidth = 4.2780 # [mm] warp width: from IMAGE ANALYSIS data
weftGap = weftSpacing - weftWidth # [mm] space in 1 warp period not
occupied

# Cross-section exponent for Power ellipse


crosssectionPower = 0.4 # from 2010 Karahan, Lomov

# Give an initial height (all res=32 unless specified)


fabThickness = 0.682 # NF=0.99959, FVC_tows_avg = 0.680209,
FVC_textile = 0.5598573

textileResolution = 32
zshift = 0.0

# Parameters for unit cell


numWarps = 2 # number of warp yarns
numWefts = 2 # number of weft yarns

# For layered textiles / nesting


numLayers = 3

elif fabricType == "TWILL":

# Fabric
arealWeight = 280.0 # [g/m^2]
fiberDensity = 2.58 # [g/cm^3] E-glass

# Warp
warpTexStr = "EC9-68x3 t0"
warpTexNum = 204.0 # [g/1000m]
warpFibDiameter = 9.0 # [um] diameter of 1 fiber filament
warpCount = 7.0 # [warps/cm]
warpSpacing = 10./warpCount # [mm] spacing between two warps in [mm]
warpWidth = 1.3203 # [mm] warp width: from IMAGE ANALYSIS data
warpGap = warpSpacing - warpWidth # [mm] space in 1 warp period not occupied

# Weft
weftTexStr = "EC9-204"
weftTexNum = 204.0 # [g/1000m]
weftFibDiameter = 9.0 # [um]
weftCount = 6.5 # [warps/cm]

    2
UnitCellGenerator.py  
   
 
 

weftSpacing = 10./weftCount # [mm] spacing between two wefts in [mm]


weftWidth = 1.5045 - 0.1 # [mm] weft width: from IMAGE ANALYSIS data
weftGap = weftSpacing - weftWidth # [mm] space in 1 weft period not occupied

# Cross-section exponent for Power ellipse


crosssectionPower = 1.8

# Give an initial height


fabThickness = 0.385 # FVC_tows_avg = 0.51214, FVC_textile = 0.40094
zshift = -0.03
textileResolution = 24

# Parameters for unit cell


numWarps = 4 # number of warp yarns
numWefts = 4 # number of weft yarns

# For layered textiles / nesting


numLayers = 1

else:
print("Invalid fabric type! Please enter either PW, BIAXIAL, or TWILL.")

# Name for TexGen and export


textileName = str(fabricType)+"_FVC_058"+"_res"+str(textileResolution)
fileName = str(fabricType)+"_FVC_058"+"_res"+str(textileResolution)

#####################
# YARN GEOMETRIES #
#####################

if fabricType == "BIAXIAL":
warpHeight = fabThickness/2.0 - 0.01 # Added to correct interference in
BIAXIAL only
else:
warpHeight = fabThickness/2.0

weftHeight = warpHeight

# Define initial Cross-sections


warpCrossSection = CSectionPowerEllipse(warpWidth, warpHeight, crosssectionPower,
0)
warpCSArea = warpCrossSection.GetArea(warpCrossSection.GetPoints(20)) # [mm^2]

weftCrossSection = CSectionPowerEllipse(weftWidth, weftHeight, crosssectionPower,


0)
weftCSArea = weftCrossSection.GetArea(weftCrossSection.GetPoints(20)) # [mm^2]

print("Initially: warp CS area = "+ str(warpCSArea) + " [mm^2]")


print("Initially: weft CS area = "+ str(weftCSArea) + " [mm^2]")

# If warp and weft Tex# are equal, then equalize their cross sections also
if (warpTexNum == weftTexNum) and (abs(warpCSArea - weftCSArea) > 0.001):

# Reduce warp OR weft height, depending on which area is larger


if warpCSArea > weftCSArea:
print("Warp and Weft CS area are not equal, reducing warp height...")
while warpCSArea > weftCSArea:
warpHeight -= 0.001

# Update cross-section and recalculate area


warpCrossSection = CSectionPowerEllipse(warpWidth, warpHeight,
crosssectionPower, 0)
warpCSArea = weftCrossSection.GetArea(warpCrossSection.GetPoints(20))
# [mm^2]
else:
print("Warp and Weft CS area are not equal, reducing weft height...")

    3
UnitCellGenerator.py  
   
 
 

while weftCSArea > warpCSArea:


weftHeight -= 0.001

# Update cross-section and recalculate area


weftCrossSection = CSectionPowerEllipse(weftWidth, weftHeight,
crosssectionPower, 0)
weftCSArea = weftCrossSection.GetArea(weftCrossSection.GetPoints(20))
# [mm^2]

print("Warp CS area = " + str(warpCSArea) + " [mm^2]")


print("Weft CS area = " + str(weftCSArea) + " [mm^2]")

print("Initial fabric parameters:")


print("Warp height = " + str(warpHeight) + " [mm]")
print("Weft height = " + str(weftHeight) + " [mm]")
print("Layer thickness = " + str(fabThickness) + " [mm]")
print("Textile resolution = " + str(textileResolution))

#####################
# UNIT CELL #
#####################

# Create a 2D weave textile unit cell


weave = CTextileWeave2D(numWarps,numWefts,warpSpacing,fabThickness, True, False)

weave.SetXYarnWidths(warpWidth)
weave.SetXYarnHeights(warpHeight)
weave.SetXYarnSpacings(warpSpacing)

weave.SetYYarnWidths(weftWidth)
weave.SetYYarnHeights(weftHeight)
weave.SetYYarnSpacings(weftSpacing)

if fabricType == "PW":
# Set the weave pattern - PLAIN WEAVE WOVEN
weave.SwapPosition(1, 0)
weave.SwapPosition(0, 1)

elif fabricType == "BIAXIAL":


# Set the weave pattern - 0/90 Non-crimped
weave.SwapPosition(0, 0)
weave.SwapPosition(0, 0)

elif fabricType == "TWILL":


# Set the weave pattern - TWILL
weave.SwapPosition(0, 0)
weave.SwapPosition(0, 3)
weave.SwapPosition(1, 2)
weave.SwapPosition(1, 3)
weave.SwapPosition(2, 1)
weave.SwapPosition(2, 2)
weave.SwapPosition(3, 0)
weave.SwapPosition(3, 1)
else:
print("Invalid fabric type! Please enter either PW, BIAXIAL, or TWILL.")

#####################
# YARNS #
#####################

# Apply cross section properties to all yarns


print("Applying yarn properties...")
# Warp Yarns
for i in range(weave.GetNumXYarns()):
yarnTuple = weave.GetXYarns(i)

    4
UnitCellGenerator.py  
   
 
 

yarn = yarnTuple[0]

# Instantiate a yarn cross section (from a CSection)


myYarnSection = CYarnSectionInterpNode(True, False, True)

# Assign a CSection to each node for the current yarn


for j in range(yarn.GetNumNodes()):
myYarnSection.AddSection(warpCrossSection)

# Add the CYarnSection to the CYarn


yarn.AssignSection(myYarnSection)

# Set the resolution of the surface mesh created (# number points around
crosssection)
yarn.SetResolution(textileResolution)

# Set the yarn fiber density and linear density


yarn.SetFibreDensity(fiberDensity, "g/cm^3")
yarn.SetYarnLinearDensity(warpTexNum, "g/km")

# Weft Yarns
for i in range(weave.GetNumYYarns()):
yarnTuple = weave.GetYYarns(i)
yarn = yarnTuple[0]

# Instantiate a yarn cross section (from a CSection)


myYarnSection = CYarnSectionInterpNode(True, False, True)

# Assign a CSection to each node for the current yarn


for j in range(yarn.GetNumNodes()):
myYarnSection.AddSection(weftCrossSection)

# Add the CYarnSection to the CYarn


yarn.AssignSection(myYarnSection)

# Set the resolution of the surface mesh created (# number points around
crosssection)
yarn.SetResolution(textileResolution)

# Set the yarn fiber density and linear density


yarn.SetFibreDensity(fiberDensity, "g/cm^3")
yarn.SetYarnLinearDensity(weftTexNum, "g/km")

#####################
# REFINE TEXTILE #
#####################

# This domain is needed for the function RefineTextile()


# Create a domain and assign it to the textile
weave.AssignDomain(CDomainPlanes(XYZ(-weftSpacing/2., -warpSpacing/2., 0),
XYZ(numWefts*weftSpacing-weftSpacing/2., numWarps*warpSpacing-warpSpacing/2.,
fabThickness)))

# Check for interference (using CTextile.DetectInterference(...))


distanceToSurface = FloatVector()
yarnIndices = IntVector()
interferingPoints = CMesh()
numInterferingPoints = weave.DetectInterference(distanceToSurface, yarnIndices,
True, interferingPoints)

if numInterferingPoints == 0:
print("No weave interference detected")
else:
print(str(numInterferingPoints) + " inteference points detected: adjusting
weave with RefineTextile()...")

# RefineTextile corrects interference based on gapSize


weave.RefineTextile()

    5
UnitCellGenerator.py  
   
 
 

numInterferingPoints = weave.DetectInterference(distanceToSurface, yarnIndices,


True, interferingPoints)
print("After calling RefineTextile(), " + str(numInterferingPoints) + "
interference points remain")

#####################
# LAYERED TEXTILE #
#####################
# Create a CTextileLayered based on the existing CTextileWeave2D

if numLayers == 1:
print("Generating a single layered textile (weave)...")

# Create a domain and assign it to the textile


weave.AssignDomain(CDomainPlanes(XYZ(-weftSpacing/2., -warpSpacing/2., 0),
XYZ(numWefts*weftSpacing-weftSpacing/2., numWarps*warpSpacing-warpSpacing/2.,
fabThickness)))

# Add the textile to the TexGen singleton


AddTextile(textileName,weave,True)

elif (numLayers > 1) and (numLayers < 10):


print("Generating a layered textile with "+str(numLayers)+" layers...")

layeredTextile = CTextileLayered()

Offset = XYZ()
for i in range(numLayers):
layeredTextile.AddLayer(weave, Offset)
Offset.z += fabThickness

#####################
# DOMAIN #
#####################
# X-direction ==> WARPS
# Y-direction ==> WEFTS

# Full domain for all 3 layers


layeredDomain = CDomainPlanes(XYZ(-weftSpacing/2., -warpSpacing/2., 0), XYZ(
numWefts*weftSpacing-weftSpacing/2. , numWarps*warpSpacing-warpSpacing/2.,
layeredTextile.GetNumLayers()*(fabThickness)))
layeredTextile.AssignDomain(layeredDomain)

# Calculate Nesting Factor


domainLowerBound = layeredTextile.GetDomain().GetMesh().GetAABB()[0]
domainUpperBound = layeredTextile.GetDomain().GetMesh().GetAABB()[1]
domainInitialHeight = domainUpperBound.z - domainLowerBound.z
print("Domain height = " + str(domainInitialHeight) + " [mm]")
print("Nesting factor = " + str( domainInitialHeight/(numLayers*fabThickness)
))

#####################
# NESTING #
#####################
# BEFORE using NestLayers(), we must have specified a domain for our
CTextileLayered
# but AFTER call AddTextile() to the TexGen singleton.
# Note that NestLayers() automatically reduces the height of the domain after
nesting

print("Nesting layers...")

# Apply offsets to layers (alternating)


# X must be shifted by WEFTspacing, Y by WARPspacing!!
Offsets = XYVector()

    6
UnitCellGenerator.py  
   
 
 

for i in range(numLayers):
if i % 2 == 0:
# Offsets.append(XY(weftSpacing/2.0+0.1, warpSpacing/2.0+0.1))
Offsets.append(XY(weftSpacing/2.0, warpSpacing/2.0))
else:
Offsets.append(XY(0, 0))

layeredTextile.SetOffsets( Offsets )
layeredTextile.NestLayers()

# Calculate Nesting Factor


domainLowerBound = layeredTextile.GetDomain().GetMesh().GetAABB()[0]
domainUpperBound = layeredTextile.GetDomain().GetMesh().GetAABB()[1]
domainNestedHeight = domainUpperBound.z - domainLowerBound.z
print("Domain height = " + str(domainNestedHeight) + " [mm]")
print("Nesting factor = " + str( domainNestedHeight/(numLayers*fabThickness) ))

domainHeightReduction = domainInitialHeight - domainNestedHeight

#####################
# INTERFERENCE #
#####################
# After NestLayers(), check for interference of nodes and correct it

interferenceCorrection = 0.010 # Height to translate layers if interference


is detected

# Check for interference (using CTextile.DetectInterference(...))


distanceToSurface = FloatVector()
yarnIndices = IntVector()
interferingPoints = CMesh()
numInterferingPoints = layeredTextile.DetectInterference(distanceToSurface,
yarnIndices, True, interferingPoints)

if numInterferingPoints == 0:
print("No interference detected")
else:
print(str(numInterferingPoints) + " inteference points detected: adjusting
layer heights by: "+str(interferenceCorrection)+"...")

layerOffset = XYZ(0,0,0)
for i in range(layeredTextile.GetNumYarns()):

# Increment offset after looping through all yarns of one weave (layer)
if (i != 0) and (i % weave.GetNumYarns() == 0):
layerOffset.z += interferenceCorrection

# print("yarn #" + str(i) + ": Offset.z = " + str(layerOffset.z))


layeredTextile.GetYarn(i).Translate(layerOffset)

# Apply interference correction to domain limits


layeredDomain = CDomainPlanes(XYZ(-weftSpacing/2., -warpSpacing/2., 0),
XYZ( numWefts*weftSpacing-weftSpacing/2. , numWarps*warpSpacing-warpSpacing/2.,
domainNestedHeight+layerOffset.z))
layeredTextile.AssignDomain(layeredDomain)

numInterferingPoints = layeredTextile.DetectInterference(distanceToSurface,
yarnIndices, True, interferingPoints)
print("After interference correction, " + str(numInterferingPoints) + "
interference points remain")

# Calculate Nesting Factor


domainLowerBound = layeredTextile.GetDomain().GetMesh().GetAABB()[0]
domainUpperBound = layeredTextile.GetDomain().GetMesh().GetAABB()[1]
domainInteferenceCorrectedHeight = domainUpperBound.z - domainLowerBound.z
print("Domain height = " + str(domainInteferenceCorrectedHeight) + " [mm]")
print("Nesting factor = " + str(
domainInteferenceCorrectedHeight/(numLayers*fabThickness) ))

    7
UnitCellGenerator.py  
   
 
 

#####################
# TRIM DOMAIN #
#####################

# Find top and bottom weft yarns, cut domain there


topWeftYarnIndex = layeredTextile.GetNumYarns() - numWefts
layeredTextile.GetYarn(2).GetNode(1).GetPosition() # Bottom
weft yarn, middle node
layeredTextile.GetYarn(topWeftYarnIndex).GetNode(1).GetPosition() # Top weft
yarn, middle node

z0 = layeredTextile.GetYarn(2).GetNode(1).GetPosition().z #
Bottom
z1 = layeredTextile.GetYarn(topWeftYarnIndex).GetNode(1).GetPosition().z #
Top

# MESHING: use parameter zshift to select a unit cell that is "easy" to mesh
(avoid slivers)
if fabricType == "BIAXIAL":
# For BIAXIAL, we exploit symmetry - domain optimized for COMSOL meshing
layeredDomain = CDomainPlanes(XYZ(weftSpacing*3/4, warpSpacing*3/4,
z0+zshift), XYZ( (numWefts*weftSpacing) + weftSpacing*3/4, (numWarps*warpSpacing)/2
+ warpSpacing*3/4, z1+zshift))
else:
layeredDomain = CDomainPlanes(XYZ(weftSpacing/2., warpSpacing/2.,
z0+zshift), XYZ( numWefts*weftSpacing+weftSpacing/2. ,
numWarps*warpSpacing+warpSpacing/2., z1+zshift))

layeredTextile.AssignDomain(layeredDomain)

# Add the textile to the TexGen singleton


AddTextile(textileName,layeredTextile,True)

else:
print("Number of layers ("+str(numLayers)+") does not make sense.")

# Output CDomain properties


print("Domain XYZ limits: " +
str(GetTextile(textileName).GetDomain().GetMesh().GetAABB()) )

#####################
# FVC Calcs #
#####################

if GetTextile(textileName).GetDomain():
# Volume of domain
Vdom = GetTextile(textileName).GetDomain().GetVolume()

# Effective stack thickness


domainHeight = GetTextile(textileName).GetDomain().GetMesh().GetAABB()[1].z -
GetTextile(textileName).GetDomain().GetMesh().GetAABB()[0].z

# Calculate tow volume for a trimmed/untrimmed domain


calcMesh = CMesh()
GetTextile(textileName).AddVolumeToMesh(calcMesh,True)
Vtows = calcMesh.CalculateVolume()

# FVC of warp and weft yarns


FVCtowsList = []
for i in range(GetTextile(textileName).GetNumYarns()):
yarn = GetTextile(textileName).GetYarn(i)
FVCtowsList.append(yarn.GetFibreYarnVolumeFraction())
if warpCSArea < weftCSArea:
FVCtow_warp = max(FVCtowsList)
FVCtow_weft = min(FVCtowsList)
else:
FVCtow_warp = min(FVCtowsList)

    8
UnitCellGenerator.py  
   
 
 

FVCtow_weft = max(FVCtowsList)

# FVC of all the tows (warp and weft may be different, this is combined
(Vfwarp+Vfweft)/(Vwarp+Vweft))
FVC_tows = GetTextile(textileName).GetFibreYarnVolumeFraction()

# FVC of the unit cell (trimmed textile)


FVC = (FVC_tows*Vtows) / Vdom

print("Domain height = " + str(domainHeight) + " [mm]")


print("Domain volume = " + str(Vdom) + " [mm^3]")
print("Tow volume = " + str(Vtows) + " [mm^3]")
print("FVC_warp_tows = " + str(FVCtow_warp))
print("FVC_weft_tows = " + str(FVCtow_weft))
print("FVC_tows_avg = " + str(FVC_tows))
print("FVC_textile = " + str(FVC))

else:
print("No domain, skipping FVC calculations.")

FVC_tows = GetTextile(textileName).GetFibreYarnVolumeFraction()
print("FVC_tows_avg = " + str(FVC_tows))

#####################
# RENDERING #
#####################

# Rendering options
GetRenderWindow(textileName).SetXRay(True)
GetRenderWindow(textileName).RenderNodes(textileName)
GetRenderWindow(textileName).RenderInterference(textileName)
GetRenderWindow(textileName).RenderDomainAxes(textileName)

# Re-render
GetRenderWindow(textileName).RemoveTextiles()
GetRenderWindow(textileName).RenderTextile(textileName)

#####################
# EXPORT #
#####################

if exportSwitchSTEP == True:
print("Exporting fabric to CAD .stp file...")
exporter = CExporter()
exporter.SetFaceted(0)
exporter.SetExportDomain(1)
exporter.SetSubtractYarns(0)
exporter.SetJoinYarns(0)
exporter.OutputTextileToSTEP(r"C:\Users\skruesi\Documents\TexGen\Scripts\\" +
fileName + ".stp", textileName)
#
exporter.OutputTextileToIGES(r"C:\Users\skruesi\Documents\TexGen\Scripts\PlainWeave
\\" + fileName + ".iges", textileName)
print("CAD export successful")

if exportSwitchMESH == True:
print("Exporting mesh to Abaqus .inp file...")
mesher = CMesher(0) # 0/1/4 = Material Continuum /
Single Layer RVE / None
mesher.SetQuadratic(False)
mesher.SetProjectMidSideNodes(False)
mesher.SetPeriodic(True)
mesher.SetSeed(0.100)
mesher.SetMergeTolerance(0.01) # Merge nodes that are below this
tol

    9
UnitCellGenerator.py  
   
 
 

mesher.CreateMesh(textileName)
mesher.SaveVolumeMeshToABAQUS(r"C:\Users\skruesi\Documents\TexGen\Scripts\\" +
fileName + ".inp", textileName)
print("Mesh export successful")

if exportSwitchPNG == True:
GetRenderWindow(textileName).RemoveNodes()
GetRenderWindow(textileName).SetXRay(False)
GetRenderWindow(textileName).RemoveTextiles()
GetRenderWindow(textileName).RenderTextile(textileName)
GetRenderWindow(textileName).SetBackgroundColor(COLOR(1.000000000000e+000,
1.000000000000e+000, 1.000000000000e+000))
GetRenderWindow().TakeScreenShot(r"C:\Users\skruesi\Documents\TexGen\Scripts\\"
+ fileName + ".png", 3)

    10
UnitCellGenerator.py  
   
Appendix C

Calculation of Pressure Gradient

In an effort to compare the simulation with the experimental results (Section 5.3), the
pressure drop over the dual-scale fabric needed to be determined.

The experiment applied a pressure difference of 15.0 kPa from pot to pot. Because of the
presence of other components (tubes, random mat fabric, empty cavities), the pressure
difference over the dual-scale fabric is less than the total pressure difference. Figure C.1
shows a schematic of the pressure-relevant components.

The pressure drop in the resin tubes is calculated using Hagen Poiseuille’s equation [70]:

dP 8µQ
= (C.1)
dx Hagen πr4

For the pressure drop in the tool cavities, the equation for flow between two parallel
plates [71] separated by a distance d is used:

dP 12µQ
= (C.2)
dx plates d3 w

The pressure gradient in the random mat fabric can be calculated using Darcy’s Law.
For the random mat, the permeability value Krnd = 0.5 × 10−9 m2 is taken from [72].

dP µQ
= (C.3)
dx Darcy Krnd A

The experiment used a flow rate of Q = 19.16 mm3 /s and tube radii of r = 2.7 mm.
Figure C.2 shows the calculated pressure drop over each segment of the experiment.

105
Appendix C: Pressure Gradient 106

t
y

ity

ity

tle
ma .

we in
vit
be

e
nd

ave
et

tub
v

v
la
t

u
l

_ca

_ca

_ca
_tu

_ra
_in

_p

_o
_
dP

dP

dP

dP

dP

dP

dP

dP

dP
Figure C.1: Schematic of RTM injection experiment

16

Pin =11.76 kPa


12
Pressure [kPa]

4
Pressure drops in RTM experiment
Pressure drop over PW fabric Pout =0.25 kPa
0
900 1000 1100 1200 1300 1400 1500
Distance [mm]

100

80
Pressure drop [%]

60

40

20

0
1 2 3 4 5 6 7
Segment [-]

Figure C.2: Calculated pressure drops over the experiment components

The plain weave fabric (segment 5) clearly accounts for the largest pressure drop in the
entire experiment: 76.7%. There is a 21.3% drop over the random mat (segment 3).
Only 2% of the applied pressure is lost to the rest of the experimental setup.

The resulting pressure drop over the plain weave fabric is 11.5 kPa. For a unit cell
of 3.08 mm length, this gives a pressure drop of 197.8 Pa. For the simulation, this is
approximated as a pressure difference of 200 Pa.
Appendix D

Datasheets

This appendix contains the manufacturer’s data sheets of the materials modelled in this
thesis. The analysis of these materials was given in Section 2.1. The data sheets are
listed in the following order:

• Glass plain weave fabric

• Glass 2/2 twill fabric

• Carbon unidirectional fabric

• Ethylcinnamate fluid

107
GLASFILAMENTGEWEBE für die KUNSTSTOFFVERSTÄRKUNG
PRODUKTSPEZIFIKATION
Prüfnorm

Qualitäts-Nummer 92115 MIL-Y-1140H


US-Style MIL-C-9084
WLB Nr. DIN 65066
British Standard BS 3396

Finish/Ausrüstung FK144

Einheit Toleranz Prüfnorm

Bindung Leinwand DIN ISO 9354


Flächengewicht g / m² 280,0 ± 5% DIN EN 12127

Garn tex DIN EN 12654


Kettgarn EC9-68x3 t0

Schußgarn EC9-204

Fadendichte Fd / cm DIN EN 1049


Kette 7,0 ± 5%
Schuß 6,5 ± 5%

Temperaturbelastung 1)
Dauerbelastung °C 260
kurzzeitige Belastung °C 600

Feuchtegehalt % < 0,2 ± 1% DIN EN 3616


Finishgehalt % 0,08 - 0,28 ± 5% DIN ISO 1887
DIN EN 60

Dicke (Richtwert trocken) mm 0,30 ± 5% DIN ISO 4603/E


im Laminat (43% Vol.) mm 0,25 ± 5%

Die o.g. Daten beschreiben technische Belange nach dem heutigen Stand unserer Kenntnisse. Sie stellen weder Qualitätsmerkmale dar noch entbinden
sie von der Eigenverantwortlichkeit beim Umgang mit INTERGLAS Technologies Geweben. Änderungen sind vorbehalten.

P-D INTERGLAS TECHNOLOGIES AG, Benzstraße 14, D-89155 Erbach, Telefon +49 (0)7305 / 955-416, Fax +49 (0)7305 / 955-512
TS / 02.07.2009 Seite 1 von 2
GLASFILAMENTGEWEBE für die KUNSTSTOFFVERSTÄRKUNG
PRODUKTSPEZIFIKATION
Prüfnorm

Qualitäts-Nummer 92125 MIL-Y-1140H


US-Style MIL-C-9084
WLB Nr. 8.4551.60 DIN 65066
British Standard BS 3396

Finish/Ausrüstung FK144

Einheit Toleranz Prüfnorm

Bindung Köper 2/2 DIN ISO 9354


Flächengewicht g / m² 280,0 ± 5% DIN EN 12127

Garn tex DIN EN 12654


Kettgarn EC9-68x3 t0

Schußgarn EC9-204

Fadendichte Fd / cm DIN EN 1049


Kette 7,0 ± 5%
Schuß 6,5 ± 5%

Temperaturbelastung 1)
Dauerbelastung °C 260
kurzzeitige Belastung °C 600

Feuchtegehalt % < 0,2 ± 1% DIN EN 3616


Finishgehalt % 0,08 - 0,28 ± 5% DIN ISO 1887
DIN EN 60

Dicke (Richtwert trocken) mm 0,35 ± 5% DIN ISO 4603/E


im Laminat (43% Vol.) mm 0,25 ± 5%

Die o.g. Daten beschreiben technische Belange nach dem heutigen Stand unserer Kenntnisse. Sie stellen weder Qualitätsmerkmale dar noch entbinden
sie von der Eigenverantwortlichkeit beim Umgang mit INTERGLAS Technologies Geweben. Änderungen sind vorbehalten.

P-D INTERGLAS TECHNOLOGIES AG, Benzstraße 14, D-89155 Erbach, Telefon +49 (0)7305 / 955-416, Fax +49 (0)7305 / 955-512
TS / 02.07.2009 Seite 1 von 2
SAERTEX GmbH & Co. KG
Brochterbecker Damm 52
48369 Saerbeck / Germany
tel phone +49-25 74-9 02 0
fax +49-25 74-9 02 209
e-Mail: info@saertex.com

DATENBLATT
(nach EN 13473-1)

SAP-MATERIAL-NR.
30002338

TEXTILE STRUKTUR
7001780
ARTIKEL-BEZEICHNUNG
U-C-PB-336g/m²-1270mm

Z
=
KONSTRUKTION FLÄCHENGEWICHT TOLERANZ MATERIAL
[g/m²] [+/-%]
Oberlage
0° 313 5 TORAY T620SC 50C 24K
90 ° 9 5 E-Glas 34 tex
Pulver 8 20 Momentive Epikote Resin 05390
< X/Y
Unterlage

NÄHFADEN: 6 g/m² +/- 1 g/m² PES 76 dtex

Nähbindung: Trikot Nähfeinheit : 5,0


Breite:: 1.270 mm
Flächengewicht: 336 g/m² Gesamttoleranz: 5,3 %
SIGMA-ALDRICH sigma-aldrich.com

SICHERHEITSDATENBLATT
gemäß Verordnung (EG) Nr. 1907/2006
Version 5.2 Überarbeitet am 05.05.2014
Druckdatum 28.01.2015

ABSCHNITT 1: Bezeichnung des Stoffs bzw. des Gemischs und des Unternehmens
1.1 Produktidentifikatoren
Produktname : Ethylcinnamat
Produktnummer : W243000
Marke : Aldrich
REACH Nr. : Eine Registriernummer für diesen Stoff ist nicht vorhanden, da der Stoff
oder seine Verwendung von der Registrierung
ausgenommen sind, die jährliche Tonnage keine Registrierung erfordert
oder die Registrierung für einen späteren
Zeitpunkt vorgesehen ist.
CAS-Nr. : 103-36-6
1.2 Relevante identifizierte Verwendungen des Stoffs oder Gemischs und Verwendungen, von denen
abgeraten wird
Identifizierte : Laborchemikalien, Herstellung von Stoffen
Verwendungen
1.3 Einzelheiten zum Lieferanten, der das Sicherheitsdatenblatt bereitstellt
Firma : Sigma-Aldrich Chemie GmbH
Industriestrasse 25
CH-9471 BUCHS
Telefon : +41 81-755-2511
Fax : +41 81-756-5449
Email-Adresse : eurtechserv@sial.com
1.4 Notrufnummer
Notfall Tel.-Nr. : +41 81-755-2255
145(CH)
+41 44-251-5151 (Tox-Zentrum)

ABSCHNITT 2: Mögliche Gefahren


2.1 Einstufung des Stoffs oder Gemischs
Kein gefährlicher Stoff oder gefährliches Gemisch gemäss der Verordnung (EG) Nr. 1272/2008.
Dieser Stoff ist gemäß Richtlinie 67/548/EWG nicht als gefährlich eingestuft.
2.2 Kennzeichnungselemente
Das Produkt ist nach EG-Richtlinien oder den jeweiligen nationalen Gesetzen nicht
kennzeichnungspflichtig.
2.3 Weitere Gefahren - kein(e,er)

ABSCHNITT 3: Zusammensetzung/Angaben zu Bestandteilen


3.1 Stoffe
Formel : C11H12O2
Molekulargewicht : 176,21 g/mol
CAS-Nr. : 103-36-6
EG-Nr. : 203-104-6

Aldrich - W243000 Seite 1 von 6

You might also like