Sei sulla pagina 1di 48

3D-woven Reinforcement in Composites

FREDRIK STIG

Doctoral Thesis
Stockholm, Sweden 2012

TRITA AVE 2012:01


ISSN 1651-7660
ISBN 978-91-7501-245-2

KTH School of Engineering Sciences


SE-100 44 Stockholm
SWEDEN

Akademisk avhandling som med tillstnd av Kungl Tekniska hgskolan framlgges


till offentlig granskning fr avlggande av teknologie doktorsxamen i lttkonstruktioner fredagen den 17 februari 2012 klockan 10.15 i sal D3, Lindstedtsvgen 5,
Kungliga Tekniska Hgskolan, Stockholm.
Fredrik Stig, 2012
Tryck: E-print

Abstract
Composites made from three-dimensional (3D) textile preforms can reduce
both the weight and manufacturing cost of advanced composite structures
within e.g. aircraft, naval vessels and blades of wind turbines. In this thesis
composite beams reinforced with 3D weave are studied, which are intended
for use as joining elements in a boltless modular design.
In practice, there are a few obstacles on the way to realise the modular
boltless design. There is lack of experimental data and more importantly, lack
of experience and tools to predict the properties of composites reinforced with
3D-weaves. The novel material will not be accepted and used in engineering
applications unless proper design methods are available.
The overall aim of this thesis is to remedy these deficiencies by generating
data, experience and a foundation for the development of adequate design
methods.
In Paper A, an initial experimental study is presented where the mechanical
properties of 3D-weave reinforced composites are compared with corresponding properties of 2D-laminates. The conclusion from Paper A is that the outof-plane properties are enhanced, while the in-plane stiffness and strength is
reduced.
In Paper B the influential crimp parameter is investigated and three analytical models are proposed. The warp yarns exhibit 3D crimp which had
a large effect the predicted Youngs modulus as expected. The three models
have different levels of detail, and the more sophisticated models generate
more reliable predictions. However, the overall trends are consistent for all
models.
A novel framework for constitutive modelling of composites reinforced with
3D-woven preforms is presented in Papers C and D. The framework enables
predictive modelling of both internal architecture and mechanical properties
of composites containing 3D textiles using a minimum of input parameters.
The result is geometry models which are near authentic with a high level of
detail in features compared with real composite specimens. The proposed
methodology is therefore the main contribution of this thesis to the field of
composite material simulation.
Paper E addresses the effect of crimp and different textile architectures on
the mechanical properties of the final composite material. Both stiffness and
strength decreases non-linearly with increasing crimp. Furthermore specimens
containing 3D-woven reinforcement exhibit non-linear stress-strain behaviour
in tension, believed to be associated with relatively early onset of matrix shear
cracks.

iii

Preface
The work presented in this thesis was carried out at the Department of Aeronautical and Vehicle Engineering, School of Engineering Sciences, Kungliga Tekniska
Hgskolan (KTH). Funding was provided by the European Commission through
the research programs MOJO (AST5-CT-2006-030871) and CERFAC (ACPO-GA2010-266026). The financial support is gratefully acknowledged.
I would like to express my deepest gratitude to my supervisor Dr. Stefan Hallstrm for his excellent guidance, tireless support and positive aura. The beauty of
our relation is that we both learn from each other. I learn the art of science, and
Stefan the art of how to hit the perfect one handed straight backhand. I am grateful for the opportunity to work with the new and exciting 3D-woven material, and
who would have thought that I could put loom builder on my resum. Dr Nandan
Khokar, the inventor of 3D-weaving, is gratefully acknowledged for introducing me
to the world of weaving.
Furthermore, many thanks to my friends and colleagues, past and present, at the
Division of Lightweight Structures for your encouragement, support and lively lunch
discussions. I would especially like to thank my office room mate and good friend
Christopher Cameron for being my partner in our ski-building project, that started
as an after-hours hobby project and ended as a course for the fourth year students.
Thanks also to Anders Lindstrm for all the good times, Marianne Jacobsen for
morale-boosting and Ylva Larberg without whom I would have no clue of whats
really going on at the department.
Finally, I would like to thank my beloved parents and sister for your endless support. Julia, my friend inspiration and love who can relate to the ups and downs of
research, and my daughter Emelie for setting all the ups and downs in perspective
with a giggle.

Fredrik Stig
Stockholm, November 2011

Dissertation
This doctoral thesis consists of an introduction to the area of research and the
following appended papers:
Paper A
Fredrik Stig and Stefan Hallstrm, Assessment of the Mechanical Properties of
a new 3D woven Fibre Composite Material, Composites Science and Technology,
69(11-12):16861692, 2009.
Paper B
Fredrik Stig and Stefan Hallstrm, Influence of Crimp on 3D-woven Fibre Reinforced Composites, submitted for publication.
Paper C
Fredrik Stig and Stefan Hallstrm, Spatial Modelling of 3D-woven Textiles, Composite Structures, DOI:10.1016/j.compstruct.2011.12.003, 2011.
Paper D
Fredrik Stig and Stefan Hallstrm, A Modelling Framework for Composites containing 3D Reinforcement, submitted for publication.
Paper E
Fredrik Stig and Stefan Hallstrm, Effects of Crimp and Textile Architecture on
the Tensile Response of Composites with 3D Reinforcement, manuscript.

vii

Parts of this thesis have also been presented as follows:


F. Stig, S. Hallstrm. An Assessment of the Mechanical Properties of 3D Woven
Fibre Composite. Book of abstracts of the 8th Seminar on Experimental Techniques
and Design in Composite Materials (ETDCM-8), 3-6 October 2007: Sardinia, Italy.
F. Stig, An Introduction to the Mechanics of 3D-Woven Fibre Reinforced Composites, 2009, Licentiate thesis, ISBN 978-91-7415-295-1.
F. Stig, S. Hallstrm. Constitutive Modelling of Composite Materials with Full
3D orientation of Reinforcement. Proceedings of the 14th European Conference on
Composite Materials (ECCM-14), 7-10 June 2010: Budapest, Hungary.
F. Stig, S. Hallstrm. Modelling of Composites containing 3D-woven Reinforcement. Proceedings of the 3rd World Conference on 3D Fabrics and Their Applications, 20-21 April 2011,Wuhan, P. R. China.
Additional work to be submitted for publication:
M.W. Tahir, F. Stig, M. kermo, S. Hallstrm, Influence of Fibrous Architecture
on the Permeability of 3D-woven Textiles - Experimental and Numerical Study, to
be submitted for publication.

viii

Contents
Preface

Dissertation

vii

I Introduction
1 Background
1.1 3D reinforcement . . .
1.2 Potential applications
1.3 Benefits of 3D-fabrics
1.4 Objective . . . . . . .

1
.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

3
4
5
7
9

2 Weaving
10
2.1 2D-Weaving . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.2 3D-Weaving . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Noobing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3 Loom development

18

4 Crimp

19

5 Modelling
23
5.1 Internal geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
5.2 Mechanical properties . . . . . . . . . . . . . . . . . . . . . . . . . . 28
6 Contribution to the field & summary of appended papers

30

7 Future work

31

Bibliography

32

ix

Part I

Introduction

Introduction

Background

Three-dimensionally woven textiles are not only beautiful, they also have the potential to change the way aircraft and other complex and weight critical structures
are built.
New challenges emerges with the increasing concern for the human foot-print
on the environment. How to facilitate future demand for safe, fast, economical
and yet less environmentally harmful transportation is such a challenge. In this
quest, weight plays a paramount role since the energy required to move an object is
largely dependant on its mass. The trend in the aerospace industry, and elsewhere,
is therefore to design for both lightweight and low cost. The leading aircraft manufacturers have, in their strive to reduce structural weight, dramatically increased
the use of fibre reinforced composite materials in their new aircraft, much due to
the high stiffness and strength to weight ratio of the material. The development
can be exemplified by the new Boeing 787 Dreamliner which is the first commercial
aircraft with both composite wings and fuselage1 , for which the combined weight
of the composite parts account for 50% of the structural weight [1]. This marks
a shift from using composite materials mainly in secondary structures (i.e. radar
domes and control surfaces etc.) to also using composites in load bearing primary
structures. However, the increased use of composite materials is currently associated with increased manufacturing costs [2]. Today the industry estimates that the
market for carbon fibre reinforced plastics (CFRP) is worth about $16 billion, and
is expected to triple within 10 years [3].
The most common manufacturing method for primary aircraft structures made
of CFRP, is pre-preg layup2 . The mechanical properties of composites made from
pre-pregs are characterised by high in-plane stiffness and strength and lower outof-plane stiffness and strength. According to Naik and Ganesh [4] the majority
of primary loads in aircraft structures are in-plane, and hence motivate the use
of pre-preg laminates. The fuselage is one example where this is true. However,
in parts such as stiffeners3 and stringers4 not all loads are in-plane, making the
pre-preg laminate less suitable. Composite materials with better through-thickness
mechanical properties are thus desired to substitute todays metal parts where such
loads occur. The subject of composite stiffeners and stringers is further discussed
in section 1.2.
1 The

main body of an aircraft


is short for pre-impregnated and indicate that the fibres are impregnated with partly
hardened resin. After the pre-preg is laid up, the composite part is cured using heat and elevated
pressure.
3 Beams for stiffening plates to prevent buckling.
4 Beams in the length wise direction of the aircraft connecting the main frames, together they
make up the skeleton on on which plates are mounted to complete the fuselage.
2 Pre-preg

Fredrik Stig

1.1

3D reinforcement

Three-dimensional (3D)-fabrics were developed in the 1970s [5], and are characterised by yarns oriented not only in-plane, but also in the through-thickness
direction resulting in higher through-thickness strength and stiffness of the final
composite material. Net-shaped 3D-textile beams of various cross sections may be
directly produced by weaving or noobing (see chapter 2). These 3D-fabrics may be
subsequently used as pre-forms5 for production of composite beams, that could potentially be used as stiffeners or stringers. The fact that the 3D-fabric is a pre-form
by itself is a major advantage compared to producing a beam with the same shape
using two-dimensional (2D)-mats, which is a time-consuming and expensive multistep process (cutting mats, shaping profile and bonding it together). Prichard [6]
states that the aerospace industry needs bending and twisting stringer-like beams
with variable cross section shape along the length, and if needed, the possibility to
directly tailor the ends of the beam to facilitate the use of fasteners. This thesis
will discuss pros and cons with the novel 3D-woven 3D-reinforcement (see chapter
2.2) which could potentially meet the requirements formulated by Prichard. The
3D-woven 3D-fabric is currently being developed by the weaving innovation company Biteam AB in collaboration with the division of Lightweight Structures at
KTH.
Why then is the 3D-woven architecture so interesting apart from the obvious
benefit of increased stiffness and strength in the out-of-plane direction? The answer
lies partly in the fabrics inherent three dimensional structure creating integrity
in three mutually perpendicular directions, partly in the fact that the pre-forms
may be produced to net shape reducing overall manufacturing cost, but also in
the versatility of the weaving process. The weaving process allows for (among
other things) the usage of different types of yarns and different weave patterns
in various parts of the pre-form, making the design space enormous. It is even
possible to vary the cross-section shape and size along the length of the pre-form
by adding and removing yarns. The latter, together with the variable yarn and
weave types, make the 3D-woven profile more like a structure than homogeneous
material. It also implies that the mechanical properties may vary in different parts
of the profile. This of course adds to the complexity but it also opens up for
new design opportunities. The 3D-woven architecture could also be interesting as
reinforcement in an anisotropic composite bulk material from which a part may be
machined. The 3D-weaving process is explained and further discussed in chapter
2.2.

5A

pre-shaped fibre structure prior to infusion. The pre-form is usually made in a separate
tool. For dry 2D fabrics: a stack of pre shaped fibre mats held together by either adhesives,
stitches or staples.

Introduction
1.2

Potential applications

There are potentially many applications for these novel composite beams containing
3D-textiles. The aerospace industry has shown an early interest and is currently
researching the potential use of these beams as joining elements, stiffeners and
stringers.
The work described in this thesis is a part of two such EU-funded aerospace
projects: Modular joints for composite aircraft components (MOJO) [7] and Cost
Effective Reinforcement of Fastener Areas in Composites (CERFAC)6 . The overall
aim of the two projects is to promote the use of composite materials in primary
load bearing aircraft structures, and thereby reduce weight and ideally also assembly
cost. Both projects target challenges when joining composite parts in load bearing
structures.
Joints in composite aircraft structures is a subject of its own and will be only
briefly discussed here. The motivation for the aerospace industry to pursue a boltless design using composite beams as joining elements is associated with the current
problems with traditional joining methods such as using bolts and rivets. These
problems effectively prevent the use of composite materials to their full potential
in aircraft design.
The problem with using bolts for joining composite plates lies in the mechanical
properties of the composite material. Unlike metals, composites are anisotropic
and have a limited ability of yielding. The latter gives rise to high stress concentrations which would otherwise be relieved by plastic deformation [8]. The problem
is exemplified by McCarthy in Fig. 1, where a maximum joint efficiency for bolted
joints of composite plates (depending on joint design) are at most 40-50% (often
less), compared to 70-80% for metal joints [9].
Strength
baseline
metal
composite

100 %
70-80 %
40-50 %

Figure 1: General joint efficiencies when joining metal and composite plates (here
illustrated schematically).

The stress concentrations around the bolts in composite plates have great implications on the design and are often the active design constraint governing the
6 http://research.cenaero.be/cerfac

Fredrik Stig

minimum plate thickness [7]. It is therefore difficult to achieve more weight savings
with the presence of bolts. This implies that the plate is over-dimensioned (and
overly heavy) around the bolted areas.
A possible solution in line with the vision of a boltless design was investigated
successfully within the MOJO project. Instead of bolts, beams made from 3D textile pre-forms with different cross sections, e.g. T, and H, were used as joining
elements to assemble plates using adhesives, see Fig. 2. An alternative to using
adhesives would be to co-cure the entire part. The results of the MOJO project

Figure 2: The MOJO joint design using to the left a profile and to the right an
H profile.

indicate that a reduction in component assembly cost of 17% compared to a rivited


structure and a weight reduction by around 50% compared to a riveted metallic
structure may be achieved [10]. The weight reduction was achieved by the elimination of fasteners and the reduction of plate thickness, previously governed by the
requirements set by the use of bolts. Cost reduction was mainly achieved by the
introduction of beams as joining elements which eliminates the use of bolts, and
thereby reduces the cost of assembly. The manufacturing cost of the joining elements depends on the manufacturing method. The use of dry textile pre-forms in
an out-of-autoclave vacuum assisted impregnation process is less expensive than using pre-preg in an autoclave7 . The weight saved will decrease the fuel consumption
and hence the operating cost for the airline operators [7].
Another possible application is stiffeners and stringers. As can be seen in Fig.
3, a commercial aircraft contains a vast number of stiffeners, stringers and frames.
According to Prichard [6] the aircraft manufacturers use metallic stiffeners and
stringers by the kilometre every day. If composites are to be a viable option to
replace metal parts, then the reinforcement pre-forms have to be produced directly
to net shape, as in the case with the 3D woven profiles.
7 High temperature and high pressure chamber for curing composites, commonly used for
curing pre-preg laminates.

Introduction

Figure 3: Fuselage barrel of Airbus long range aircraft with visible stringers and
frames. Reprinted with permission from P. Linde, Airbus (Hamburg) [11].

1.3

Benefits of 3D-fabrics

As discussed, composites reinforced with net-shaped three dimensional (3D)-fabric


pre-forms have emerged as a viable solution for parts like stiffeners and stringers.
That 3D-fabric reinforcement can overcome the problems of low inter-laminar and
through-thickness strength typical for traditional 2D-laminates has been known
for many years, and was pointed out by Kregers and Melbardis [12] already in
1978. Today, the motivation for introducing 3D-textiles in composites is not only
to increase the out-of-plane properites, but also to reduce the labour cost during
manufacturing and to integrate more functionality into the component [13]. The
latter may be achieved by for instance incorporating optical fibres or various types
of wiring.
The composite beams used in the aerospace industry today are often produced by
cutting, stacking and draping 2D pre-preg laminas, see Fig. 4. There are problems
associated with this methodology, the main ones being difficulty to fixate the layers to the desired shape before curing, and to achieve sufficient through-thickness
mechanical properties of the cured parts. For instance, when manufacturing Tprofiles as illustrated in Fig. 4, limitations in curvature of the layers generate a
region at the junction that needs certain attention. The current approach is to
fill this pocket with a noodle (or gusset filler) to avoid problems associated with
resin rich areas. Nevertheless, failure in such beams usually initiates in or near
the noodle due to strong multi-axial stress fields (including peel stresses) caused
by the transfer of loads in the curved areas [7]. To improve the through-thickness

Fredrik Stig

properties, various forms of stitching, bonding or z-pinning8 have been developed,


each technique having different benefits and drawbacks [14]. One of the remedies
proposed in the MOJO project is to use 3D-woven preforms, which offer structural
integrity at handling before and during composite manufacturing. Another option
is to use e.g. 3D textile braids.

Laminas
Crack

Noodle

Figure 4: A conventional T-beam built up from 2D pre-preg laminas, loaded until


failure. Redrawn from [7].

There is however a price to be paid for higher out-of-plane properties and that
is a degradation of in-plane properties. The presence of fibres in the out-of-plane
direction not only decreases the in-plane fibre volume fraction, it also induces crimp
in the in-plane yarns. The stiffness, and more importantly also the load carrying
capacity, of yarns are significantly reduced by crimp, see chapter 4. How to find
the balance between desired improvement of the out-of-plane properties and corresponding unwanted decrease in in-plane properties is an interesting topic for future
analysis and optimisation. New and reliable modelling methods are needed to guide
engineers in this compromise.
According to McHugh [13] the future for 3D-textiles is bright. However, new and
better tools for predicting their properties are needed for these novel materials to
be accepted and used in engineering applications [13, 15].
The reinforcement architecture of a composite material plays a paramount role
for its mechanical properties [16, 17]. The properties of various other 3D-textile
reinforced composites have been extensively studied e.g. [1829]. However, the
properties of true 3D-woven reinforced composite materials are still uncharted.
Furthermore, many of the modelling methods used for the other 3D-fabric reinforced
composites are either not applicable or available for the 3D-weaves addressed in this
thesis.
8A

form a stapling with small carbon rods

Introduction
1.4

Objective

The objective of this thesis is to characterise the material and to develop the necessary modelling tools to predict and tailor the mechanical properties, and thereby
facilitate the materials future use. The thesis consists of five appended papers with
the following objectives:
Paper A To perform an initial experimental study on the mechanical properties of 3D-reinforced composites and compare them with corresponding properties of 2D-laminates.
Paper B To explore and model how three dimensional yarn crimp influences
the longitudinal Youngs modulus. New analytical models are developed.
Paper C To address the important issue of creating geometrically representative models of the 3D-textiles internal structure.
Paper D To create a framework for finite element (FE) analyses of 3D-textile
reinforced composites using geometries generated by the scheme developed
and presented in Paper C.
Paper E To experimentally determine the effects of weave architecture and
crimp on both stiffness and strength of 3D-weave reinforced composites.

10

Fredrik Stig

Weaving

Textile manufacturing processes are broadly grouped into four types - weaving,
braiding, knitting and non-wovens [30]. Each of these processes is further subdivided according to certain processing characteristics. It is possible to manufacture
3D-fabrics using all of these techniques. However, this chapter focuses on weaving,
and the aim is to give a brief overview of the two fundamentally different weaving methods (2D and 3D weaving) employed to produce woven 3D-fabrics. In the
composite materials industry today several types of weave structures are labelled
3D-woven. Khokar [3133] suggests a more distinct definition of 3D textiles distinguishing between 2D-woven 3D-fabrics (i.e. produced by conventional 2D-weaving
process), and 3D-woven 3D-fabrics (i.e. produced by 3D-weaving process) and a
new class of non-woven fabrics named noobed fabrics (produced by the non-woven
noobing process). These three processes and their products are categorised in Fig.
5 and further discussed in the following sections.
2Dweaving

3Dweaving

2Dfabrics

3Dfabrics

Nonwoven

Noobed
fabrics

3Dfabrics

Figure 5: Categorisation of 3D-fabrics and their manufacturing processes.

The fundamental operations constituting the weaving process are, in their sequential order, shedding, picking, beating up and taking up. The shedding operation
displaces the warp yarns using healds9 to create a shed10 . The various shedding11
methods available have been described by Khokar [34]. The shedding operation is
followed by the picking operation whereby the weft yarns are inserted in the created
shed. The weft that is laid in the shed is beaten-up by the reed12 to the fabric-fell
position, and thereby completing the fabric formation. To achieve continuity in the
process, the produced fabric is advanced forward by the take-up operation. This
cycle of operations are continued repeatedly to obtain the woven fabric.
In a 2D-weaving process, two mutually perpendicular sets of yarns, the warps
and the wefts, are used. The warp yarns run in the length-wise direction of the
9A

heald is a flat steel strip or a wire, with one or more eyes in which warp yarns are threaded.
gap between the two lines of warp threads caused by raising one portion and lowering the
other using healds.
11 Operation to crate a shed.
12 A metal comb fixed in the loom used for beating-up. The closeness of its teeth determines
the fineness of a cloth.
10 A

Introduction

11

fabric, and the weft yarns in the transverse direction. This arrangement remains
unchanged whether a single warp sheet is used (to produce sheet-like 2D-fabrics) or
multiple warp sheets are used (to produce 3D-fabrics). This is because the weaving
process continues to be identical in the production of both these fabric types.
In a 3D-weaving process, a grid-like set of warp yarns and two mutually perpendicular sets of weft yarns are used. These two sets of weft yarns are denoted
horizontal wefts and vertical wefts.
In both 2D- and 3D-weaving processes, additional non-interlacing stuffer warp
yarns can be included. These stuffer warp yarns do not interlace with the wefts
and hence occur linearly in the woven fabric.
A variety of basic 3D-fabric architectures are produceable employing the 2D- and
3D-weaving processes, depending on how the warp is set up and how the sheds are
formed. According to Whitney and Chou [35] the exact geometry of each of the
basic woven architectures is influenced by several weaving parameters such as:
Fibre or tow size
Number of warp and weft yarns per unit width and length
Number of warp layers interlacing with each weft yarn (weave pattern)
Tension in the warp and weft yarns
Tightness of the tows (resistance of deforming from its original cross-sectional
shape).
2.1

2D-Weaving

Two-dimensional weaving may be utilised to produce both conventional sheet-like


2D-fabrics and some 3D-fabrics. Two-dimensional weaving is characterised by
mono-directional shedding, and the interlacing of two orthogonal sets of yarns.
Khokar [31] defines the 2D-weaving process as
the action of interlacing either a single- or a multiple-layer warp with a
set of weft.
The manner in which the 2D-weaving process is employed for producing both 2Dfabrics and 3D-fabrics is illustrated in Fig. 6.
The two most common 2D-woven 3D-fabrics are the layer-to-layer angle interlock
weave and through-the-thickness angle interlock weave (also known as 3-X weave),
see Fig. 7. The 2D-woven 3D-frabrics are usually produced as wide sheets for
use in manufacturing conveyor belts, paper clothing, double cloth13 etc. The 2Dweaving technique also allows indirect production of a few, relatively simple, shell
13 Double cloth contain two or more sets of warps and one or more sets of horizontal wefts that
interlace in two-layers, allowing complex patterns and surface textures to be created.

12

Fredrik Stig

type profiled 3D-fabrics. Using multi-eyed healds in a conventional 2D-weaving


equipment, Chiu and Cheng [16] developed a method for weaving profiled materials
having double cross- (++) and I-shaped cross-sections. Such pre-forms are woven in
a flat condition and the flanges are subsequently unfolded to form a near-net-shaped
pre-form.
warp
Picking

weft

weft

Shedding
warp

warp
Healds

(a)

(b)

Figure 6: Illustration of the 2D-weaving principle for a) 2D-fabrics and b) 3Dfabrics, redrawn from [31].

x
y

Stuffer

Warp

Weft

Figure 7: A Through-the-thickness angle interlock weave with stuffer yarns (blue)


seen from its three principle planes and from an isometric view. Warp yarns are
green and weft yarns red in the illustrations.

Introduction
2.2

13

3D-Weaving

While 2D-weaving has existed for at least five millennia, 3D-weaving is a relatively
new weaving development, invented by Khokar in 1997 [32]. The 3D-weaving process is characterised by the incorporation of the dual directional shedding operation.
Such a shedding system enables the warp yarns to interlace with both horizontal
and vertical sets of weft yarns. Khokar [31] defines the 3D-weaving process as
the action of interlacing a grid-like multiple-layer warp with the sets of
vertical and horizontal wefts.
Accordingly, the warp yarns, the horizontal and vertical sets of weft yarns occur
mutually perpendicular to each other. The schematic of the 3D-weaving process is
illustrated in Fig. 8(a-h) for a plain 3D-weave, and the sub steps are explained as
follows:
(a) The warp yarns are arranged and supplied in a grid-like arrangement.
(b) Warps are displaced in the vertical direction to create multiple horizontal
sheds.
(c) Corresponding number of horizontal wefts are inserted into the created sheds.
(d) The multiple horizontal sheds are then closed, whereby the warp yarns become
interlaced with the horizontal weft yarns.
(e) The warps are at their level positions in the weaving cycle.
(f) Next, the warp yarns are displaced in the horizontal direction to create multiple vertical sheds.
(g) Corresponding number of vertical wefts are inserted into the created sheds.
(h) The multiple vertical sheds are then closed, whereby the warp yarns become
interlaced with the vertical weft yarns.
These sequences of operations are repeated once more to insert the wefts in the
respective opposite directions to complete one cycle of the 3D-weaving process to
obtain the woven structure in Fig. 8i.
An important feature of such a woven structure is the occurrence of pockets
between any four adjacent warp yarns. These pockets can be directly filled with
stuffer warp yarns (Fig. 8j) during the weaving process, although they do not
interlace with the wefts. It is not necessary to fill all the pockets; select pockets can
be filled with stuffer warp yarns according to needs. Furthermore, these pockets
could as well incorporate e.g. electrical wires, tubes, optical fibres etc.
To complement the weave illustration given in Fig. 8i, the views from the three
principle planes and from an isometric view are shown in Fig. 9. If the textile

14

Fredrik Stig
Horizontal Picking

Vertical Shedding

Warp

(a)

(b)

(c)

Horizontal Shedding

(e)

Vertical Picking

(g)

(f)

Interlacing

(d)
Interlacing

(h)

Stuffer yarns

z
x

(i)

(j)

Figure 8: Illustrations of the 3D-weaving principle, seen from the warp direction
here denoted x-direction.

architecture of such a 3D-woven 3D-fabric is compared with that of the 2D-woven


3D-fabrics illustrated in Fig. 7, a significant difference can be observed. The warp
yarns in a plain 3D-weave are fully interlaced with both the horizontal- and verticalweft yarns. In the interlock weave there is no interlacing in the y-direction, which
is a direct consequence of the fact that no horizontal shedding occurs. This is
highlighted in the x-y plane illustrations in Figs. 7 and 9. Furthermore, the warp
yarns in the 3D-woven 3D-fabric do not traverse neither between layers of the
fabric nor from one surface of the fabric to the opposite, but remain in their assigned
planes. In the 2D-woven 3D-fabrics the warp yarns traverse between layers of the
fabric (layer-to-layer angle interlock weave) and from top and bottom surfaces, as
can be observed in Fig. 7.
The 3D-weaving process allows direct production of woven profiled materials.
Such profiled materials can be produced in solid, shell, tubular and their combination types directly. Fig. 10 exemplifies cross-sections of different profiled materials
obtained directly by the 3D-weaving process. It also allows flanges with tapered
surfaces. The 3D-weaving process also offers great flexibility in creating different
weave patterns, such as plain, twill and satin, in different locations of the crosssection and along the length of the profiled pre-form. Yet another possibility is the
production of curved profiled pre-forms.

Introduction

15

z
x

x
y

Warp

V-weft

H-weft

Figure 9: A plain 3D-weave without stuffer yarns seen from its three principle
planes and from an isometric view. Warp yarns are blue, horizontal weft yarns red
and vertical weft yarns green in the illustrations.

2.3

Noobing

Noobing is a non-woven 3D fabric-forming process that essentially assembles three


mutually perpendicular sets of yarns. Its principle fabric structure is illustrated in
Fig. 11. There is no interlacing (as with weaving), interlooping (as with knitting)
or intertwining (as with braiding) of the involved yarns. This process was defined
by Khokar [31] as
the action of producing a non-woven 3D-fabric by orientating orthogonally and binding the employed essentially three sets of yarns.
The term noobing is an acronym for Non-interlacing, Orientating Orthogonally and
Binding, the main characteristics of the process and the fabric, which is called the
noobed fabric. The fabric structure is also known under a variety of different other
names. The most commonly used name is 3D orthogonal weave [23, 3639], but
there are also other names for example: XYZ fabric, zero-crimp fabric, Cartesian
fabric, DOS (directionally oriented structures) and polar fabric [31].

!"#$%&'()*+,%-%.'/)0.)1*2)3445)Fredrik Stig

16

!"#$%"$#&' ()&*+ , !-'./

;-/$'&# <-.9"+

:-#9)#

5.

6$&/#-

1)&=> <-.9.9? !>+")*

2#.

;$'".

72

09,1.9)

78

0 , ()&*

2#.-

1 , ()&*

!"#$%"$#&' ()&*+ , 2$@$'&#

49. , 2$@)

A , .9 , B

C , .9 , B

2 , ()&*

4 , ()&*

3 , ()&*

!"#$%"$#&' ()&*+ , !E)''

:-9%)9"#.%

:-9D-.9)/

7 , ()&*

F-$@') G'&9?)
(-H ()&*

Figure 10: Examples of possible cross-sections (in the y-z plane) of dry 3D-woven
pre-forms. Courtesy of Biteam AB, Sweden, www.biteam.com.
!"#$%&'(%)*+& $", $&)-'./&("/-'
-)0&'1*2"'3"-."412& -)2&/1)%'$5'
6789':;&0&*
!""#$%&#' #( )*+,-%$!". &-/0"#1#.!-2

In absence of interlacing, interlooping and intertwining of the involved yarns, the


constituent yarns occur linearly (without crimp) in their respective directions. As
a consequence, the fabrics structural integrity is achieved through bindings that
occur on the fabrics surfaces (excluding the end surfaces).
The noobing process is of two types: uniaxial and multiaxial. In the former
process a set of grid-like arranged axial yarns are bound in the fabric-width and
fabricthickness directions. In the latter process four sets of yarns are arranged in
fabric-length, fabric-width and +/- bias directions and then stitched in the fabricthickness direction. The respective 3D-fabrics are thus denoted uniaxial noobed
fabrics and multiaxial noobed fabrics.
The uniaxial noobing process has limited profiling capability only a few solid
types can be obtained (I, T, L etc. [38,40,41]). Mohamed et al. [40] state that since
there is no interlacing between the axial yarns and the binder yarns, the yarn layers
are free to shear with respect to one another and also bend making the drapability
good. The multiaxial noobing process has been successfully commercially employed
to produce sheet-like multiaxial non-crimp fabrics.

Introduction

17

z
Filler

x
y

Stuffer

Binder

Figure 11: A noobed fabric seen from its three principle planes and from an isometric view. Warp (or binder) yarns are green, stuffer yarns blue and filler yarns
red in the illustrations.

18

Fredrik Stig

Loom development

A substantial part of the first two years of working on this thesis was utilised to
validate, assemble and make operational the worlds first large scale automated 3Dweaving loom, see Fig. 12. The loom has since then been used to weave net-shaped
pre-forms for the MOJO and CERFAC projects, as well as all woven pre-forms
utilised in papers B-E.

Figure 12: The MOJO consortium members in front of the 3D-weaving loom at
KTH.

Introduction

19

Crimp

Yarn undulation, or crimp, is one of the most important properties of textiles used
as reinforcement in composite materials since it has a strong influence on the final
mechanical properties. Crimp is the topic of both Paper B and E, and therefore a
short background to the subject is given here.
There is no unified definition of crimp. Historically, crimp has had many definitions in the textile industry, and the interested reader is referred to Alexander et
al. [42] for a good summary. A common definition of crimp in the textile industry
is
yarn length
.
(1)
crimp =
wave length
Pierce [43] used this definition in one of the pioneering works from the 1930s on
textile geometry. There are also a variety of crimp definitions within the textile
composite community. The composite industry has traditionally only focused on
2D crimp, and the most common definition is the crimp ratio (CR) defined as
CR =

yarn path amplitude


.
wave length

(2)

One problem with using the CR for 3D crimp is that it yields two different ratios.
One in the x-y and one in the x-z plane according to the axis definition in Fig. 9.
Other definitions are also found in the literature, such as crimp angle [44] and an
alternative definition of the CR [45]. However, in this thesis Pierces [43] definition
of three dimensional crimp is used, since it is on a more general form and results
in only one crimp measure. Both Pierces definition of crimp and the CR are
illustrated in Fig. 13.
Ya

n g

rn

th

Amplitude

L e

Wave length

(a)

Wave length
(b)

Figure 13: The difference between the two crimp definitions. a) Illustration of the
variables in a) Eq. (1), and b) the variables in Eq. (2).

The stiffness of a composite is reduced with increasing crimp [46, 47]. This is
true also for 3D crimp and further discussed in Paper B and E. The reason is

20

Fredrik Stig

that the crimped strands14 tend to straighten out instead of stretch when loaded
in tension. To quantify the stiffness reduction, a unidirectional (UD) composite is
used as an example. Two simple models are used, both based on classical laminate
theory (CLT) [48], but with different strand path idealisations; a sine function and
a zig-zag pattern (). Two dimensional crimp is assumed and the normalised
longitudinal stiffness as function of crimp for the two models are seen in Fig. 14a
together with results from FE calculations by Garnich and Karami [47]. The figure
shows that only half of the stiffness remains when the yarns exhibit crimp of about
1.03 (illustrated for a sinusoidal yarn path in Fig. 14b).
1

0.3

0.6

0.2

0.2
0
1

Crimp = 1.01

0.1

0.4
y!axel

Normalised E [!]

0.8

Sine shape
Zig!Zag shape
Garnich and Karami
1.01

1.02

1.03
1.04
Crimp [!]

0
!0.1

Crimp = 1.0

Crimp = 1.03

!0.2

1.05

!0.3

1.06
0

0.2

0.4

0.6

0.8

x!axel

(a)

(b)

Figure 14: a) The normalised longitudinal Youngs modulus as function of crimp


for UD-composites and b) the sinusoidal yarn path idealisation for three different
crimp levels.

Crimp also affects the strength of a composite material [29,46]. Gu and Zhili [17]
argue that if all strands in a fibre reinforced composite are straight, then they
will react simultaneously to an external load and the full strength is achieved.
However, if the strands are bent, and bent with different angles, then a portion
of the strands will take up higher load and subsequently fail earlier, leading to
lower material strength. The strength as function of warp yarn crimp is studied in
Paper E, and the result can be seen in Fig. 15. West and Adams [44] quantified
the strength reduction of a 2D triaxially braided composite. Two different types
of samples were manufactured, one where the axial yarns were allowed to crimp
14 In

this thesis a distinction is made between yarn and strand. A yarn is defined as a bundle of
dry fibres, while a strand is considered as an impregnated composite yarn; hence a strand consists
of both matrix and a large number of fibres.

Introduction

21
1600
1400
Noobed

!ult [MPa]

1200
1000
800
600
400
200
0

1.01

1.02 1.03 1.04


Warp strand crimp

1.05

1.06

Figure 15: The ultimate stress of 3D-weave reinforced composites with varying warp
strand crimp. Results from a noobed fabric reinforced composite with nominally
straight stuffer yarns is added for reference.

and one where the axial yarns were kept under tension during consolidation. Axial
compression tests showed that crimp reduced the axial compression strength by
30%. DeCarvalho et al. [49] performed an experimental study on compression failure
of 2D-woven composites. They argue that the strands behave as structural elements
at meso-scale15 and that the damage behaviour in compression is dependent on the
reinforcement architecture. DeCarvalho et al. also conclude that the strands mainly
fail where the strands are crimped, and that the failure mode is predominantly
kinking16 . Budiansky and Fleck [50] presented a theoretical background of kinking
and concluded that local fibre misalignment plays an important role in the kink
band formation process.
Crimp also introduces shear stresses in the matrix [47]. Bogetti et al. [46] developed a model for predicting stiffnesses and strengths in composite laminates with
ply waviness under compressive loading. They concluded that the failure mode for
small crimp values is fibre fracture, but at a critical yarn path amplitude, interlaminar shear failure becomes the dominant failure mechanism. Cox et al. [22] studied
the strength of 2D-woven 3D-textile reinforced composites. They state that shear
stresses within crimped strands cause micro cracks. These cracks are then associ15 Three scales are used in composite mechanics: Macro, meso and micro. A composite part is
on the macro scale. The meso scale is an intermediate scale where the internal structure of the
reinforcement is described using tows or strands in a unit cell. The micro-scale is the most detailed
scale, where the interaction between matrix and individual fibres in a strand may be investigated.
16 Kinking is a micro buckling process where aligned fibres fail in band-like formation under
compressive loading.

22

Fredrik Stig

ated with a non-linearity in the stress strain curve denoted the onset of plastic tow
deformation. This non-linearity is also seen for 3D-weave reinforced composites and
further discussed in Paper E.
The crimp is dependent on yarn type, weave type and various weaving parameters
such as yarn tension in the loom. Crimp ranging between 1.01 -1.06 was measured
for the 3D weaves in Paper B. Stuffer and filler17 yarns are not part of the weave
and occur relatively straight in the preform. Lomov et al. [51] reported that stuffer
and filler crimp was below 1.001 for a noobed fabric.
To conclude, crimp is not favourable in a composite, since it degrades both stiffness and strength. On the other hand it is not possible to weave without crimp,
and crimp is associated with yarn interlacing which may be favourable for the
out-of-plane properties, for instance at composite joints.

17 Filler yarns are straight yarns perpendicular to the weaving direction but are not a part of
the weaving process, hence their relative straightness. They occur for instance in noobed fabrics.

Introduction

23

Modelling

A fibre reinforced composite is by nature heterogeneous on a micro scale and the


mechanical properties of the material are governed by its microstructure. It is,
however, impractical to also model the microstructure in structural engineering
models. Instead, homogenised anisotropic material properties are used. The challenge is to determine these properties, which usually is done on a meso-scale. The
more complex the internal geometry is, the more difficult the task will be. For a
simple cross-ply laminate CLT yields sufficiently accurate results, but as soon as
woven fabrics are considered, other more sophisticated models are needed.
5.1

Internal geometry

When the reinforcement is woven, the actual internal geometry of the composite
material is almost always complex, even if the weave pattern is simple. The strand
paths are irregular and the strand cross sections vary in both shape and size along
their trajectories. Furthermore, two strands never look exactly the same, it is
therefore necessary to generalise and introduce simplifications when describing the
internal geometry.
In spite of the difficulty, a detailed and accurate geometrical description of the
internal geometry of a textile reinforced composite is of great importance when
modelling, regardless of subsequent model strategy [52]. This is especially true if
the local stress and strain fields are of interest [53].
A model of the internal architecture may be constructed in different ways, for
instance as a computer aided design (CAD)-model, a FE mesh or by using parameters in an analytical model [54] or in a geometry pre-processor (e.g. TexGen [55]
or WiseTex [52, 56, 57]).
The model may be created in two principally different ways; experimentally or
using a modelling approach. In the former, a composite sample is first scanned and
then key textile geometry parameters are extracted and used as input in a geometry
pre-prosessor or CAD software. Alternatively, if an analytical geometry model is
used, as in Paper B or in e.g. [53, 5862], the known geometry may be translated
into various model parameters. In the modelling approach, either an analytical or
a numerical model is used to calculate the geometry. The advantages are e.g. that
the time consuming manufacturing step and the cumbersome scanning process can
be eliminated. The predictive capabilities are also increased.
There are a number of questions and difficulties to address regardless of how
the geometry is to be modelled, such as the level of detail needed and how (if
required) to avoid volume overlap in the model. The level of detail is associated
with how coarse simplifications are made. The strand path is often restricted to be
described by sine functions or by coupled polynomials (splines). The strand cross

24

Fredrik Stig

section is frequently assumed to have a circular, elliptic or lenticular18 shape and


often the shape or size is not allowed to vary over the strand length. One of the
most difficult issues when creating a solid model is to avoid volume overlap e.g.
parts of two strands occupying the same space. It originates from the fact that
the simplifications described above are too coarse, and are not able to represent
the changing strand geometries when two strands interlace. There are a number of
proposed solutions to this problem in the literature, and a new one is proposed in
Paper C.

Experimental approach to determine the internal architecture


When a composite sample is available, the internal architecture may be determined
using either a scanning electron microscope (SEM), an optical microscope, or a
micro computer tomography (CT) system [63].
The first step when using either an optical or a scanning electron microscope is
to polish the surface of the composite sample to a very fine surface finish. Then
the surface is photographed through the used microscope. To generate a 3D model,
a sequence of photographs through the depth are needed. Therefore the surface
is stepwise ground off between each consecutive photograph and a new surface is
polished and photographed, see Fig. 16. The strands cross-section shapes are found
directly from the photographs, while the strand trajectories may be approximated
by splines through their respective cross section centroids.

Figure 16: Example of how the internal geometry may be extracted by using an
optical microscope. The photographed surfaces shown are from a composite sample
reinforced with a 3D-woven preform used in Paper A.

18 shaped

like a lentil or biconvex lens

Introduction

25

With the micro CT-scan technique, a large number of two-dimensional X-ray


images are taken around a single axis of rotation. The images are then used to
create a 3D-image of the material sample, see Fig. 17a which may be sliced fictively
in any plane creating view cuts as exemplified in Fig. 17b. The benefits with the
CT-scan method over the microscope method is that it is non-destructive, and no
time consuming sectioning and polishing have to be done. The contrast in an Xray image is generated by the differences in density between the matrix and the
fibres. The density difference between glass fibres and a polymer matrix is greater
than for carbon fibres and a polymer matrix, and therefor a higher contrast in the
images is achieved. The use of micro CT-scan to extract the textile architecture
of 3D-fabric reinforced composites was validated by Desplentere et al. [64], and it
is used in Paper C to validate the geometry modelling. The dry reference weave
used for validation consisted of mainly carbon yarns and in addition a few glass
fibre tracer yarns with higher density, visible in blue in Fig. 17a and in white in
Fig. 17b. Properties such as the yarn crimp and the cross section area of the fibre
bundles, could then be extracted using image processing of the view cuts.

(a)

(b)

Figure 17: (a) A 3D-image of the dry weave used as reference in Paper C, and (b)
a view of the same weave cut perpendicular to the warp yarn direction. Visible
are also the glass fibre tracer yarns used facilitate to image post processing. Both
images are presented by courtesy of the German Aerospace Centre (DLR).

After the textile geometry is characterised by scanning, the geometry may be implemented in a geometry pre-processor or as parameters in an analytical model. For
Papers B-E, the textile geometry of the 3D-woven preforms were obtained by using
the micro CT-scan technique complemented by optical microscope measurements.

26

Fredrik Stig

Modelling approach to determine the internal architecture


A perhaps more cost-effective way to obtain the internal geometry, compared to
the experimental approach, is to use analytical or numerical modelling methods.
There are a number of such schemes proposed in the literature (e.g. [52,58,6567]).
The principle of minimum energy can be used in analytical models to calculate
the fabric geometry, as done e.g. by Sagar et al. [66] and in the textile pre-processor
WiseTex [52]. The fabric is first described using idealised shapes. Loads are then
applied and the final geometry is obtained from the state when the strain energy
in the system has a global minimum.
WiseTex allows for export of the geometry as a FE-model, but penetrating volumes are not automatically avoided. Lomov et al. [56] have therefore presented
a scheme to adjust the FE-mesh by performing an intermediate FE-calculation.
The overlapping strands are locally separated and beam elements added to provide
a small clearance between the strands, after which they are compressed together
again. The beam elements then prevent the strand volumes to penetrate each other.
Durville [68] utilised an FE-approach and modelled the filaments in each yarn
using up to 400 higher order beam elements per yarn. In the starting condition
all filaments are arranged in the same plane, and with an obvious filament volume
overlap. In the analysis step, the filaments are moved by cleaver boundary conditions and a predefined weave pattern until the volume overlaps disappear in the
dry weave.
Sherburn et. al [65] proposed an FE-based method where the yarns are represented as plates derived from the yarn mid-surfaces in a TexGen-model. The final
fabric geometry is calculated using the minimum strain energy of the plates. This
method is more suitable for 2D-textiles due to the meshing procedure where the
strands are projected onto a single plane.
Mahadik and Hallett [67] developed an algorithm for generating the internal
geometry of a layer-to-layer angle interlock weave after compaction in a mould.
The wavelength of the weave is first reduced to create a "looser" weave, and in a
second step the a tensile force was applied to the weave using a temperature drop
in an explicit FE simulation using general contact. A rigid plate was finally used
to compact the weave further.
To the authors knowledge, only Durvilles method may be applicable to model
a 3D-woven textile. This was the motivation to develop a new geometry modelling
approach presented in Paper C. In the proposed method the volumetric packing
problem is solved by modelling the strands as inflatable tubes. In the initial state
the tubes cross section areas (and thus their volumes) are made small enough
not to cause any volume interpenetrations (time T 0 in Figs. 18 and 19). In the
subsequent analysis steps, the tubes are inflated under general contact conditions
until the desired degree of packing is achieved (times T 1-T 3 in Figs. 18 and 19).
The tubes may be expanded further until near full packing is reached (see time T 4
in Fig. 19).

Introduction

27

T0

T1

T2

T3

Figure 18: The geometry simulation at four different times T0-T3. The simulation
starts at T0 and stops at T3 when the desired volume fraction of strand is reached.
The times T0-T3 are also illustrated in Fig. 19.
Modelling
results

Degree of
packing

CT-scan
image

"Fully packed"
Corresponding to
specimens used in
papers C-E

Start of spatial
simulation
T0

T1

T2

T3

T4

Simulation time

Figure 19: The degree of packing of the strands/tubes as function of simulation


time. Also visible are view cuts from the model at three time-steps and a corresponding CT-scan image. All view cuts are perpendicular to the warp strands

28

Fredrik Stig

5.2

Mechanical properties

Once the internal geometry has been established and a good geometry model is
created, the mechanical properties of the composite material can be derived. According to Tong et al. [5] and Lomov et al. [56] different modelling strategies may
be employed.
Orienting Averaging models: The material is divided into small sub-volumes
which are usually treated as UD-laminas. The homogenised properties are
computed by averaging the sub-volumes stiffness matrices using iso-strain
and/or iso-stress assumptions. For the iso-strain assumption, a prescribed
homogeneous displacement boundary condition is used implying equal strain
in all building blocks. A perfect block interfacial bond is also assumed [5].
The homogenised elastic stiffness matrix C may then be computed as
X
C=
v i Ci
(3)
i

where vi is the volume fraction and Ci the stiffness matrix of block i. For
the iso-stress assumption, the traction (applied force) is prescribed, implying
constant stress in the assembled building blocks. The homogenised elastic
stiffness matrix is subsequently computed as
X vi
1
=
.
(4)
C
Ci
i
Cox et al. [20] used the orienting averaging strategy to model a composite
reinforced with a 2D-woven 3D-fabric. Tan et al. [62] presented three mixed
iso-strain and iso-stress models for noobed fabric reinforced materials. The
analytical models developed in Paper B are also of mixed iso-strain and isostress type.
Inclusion models: The inclusion models are based on randomly dispersed
inclusions in a matrix based upon Eshelbys equivalent inclusion theory. A
good derivation of Eshelbys formula may be found in [69]. Lomov et al. [56]
point out that it is a more generalised form of an Orienting Averaging model,
but without the assumption of the same internal average stress in all phases
(or blocks). Mori and Tanaka [70] developed the theory on how to calculate
the average stresses in each phase. Gommers et al. [71] used the Mori Tanaka
theory and applied it to textile composites. One of the models in Paper B
utilises an inclusion model.
Cell models: The cell model is a FE-type model, where the internal geometry
is divided up into a regular grid of cells sometimes called voxels [72, 73] .
Finite element models: In a FE-model the material is divided into a large
number of small elements (usually referred to as a mesh), and the governing

Introduction

29

force equilibrium equations for all elements are solved simultaneously using
e.g. the principle of virtual work. The FE-approach allows for an arbitrary
description of the geometry, and if the geometry is accurate and the mesh is
refined sufficiently a detailed stress-strain state may be calculated.
The problem lies in creating good FE-models of composites reinforced with
complex textiles (e.g. 3D-weaves, 3D-braids or noobed fabrics) [73]. As previously mentioned, the models in Paper C and D both utilise the FE-approach
and address this issue.

30

Fredrik Stig

Contribution to the field & summary of appended papers

The concept of 3D-weaving is now over 15 years old, however the use of 3D-woven
pre-forms as reinforcements in composite materials is new. For the concept of
composites reinforced with 3D-woven preforms to gain acceptance as an engineering
material two areas must be addressed: the lack of analysis tools and practical
knowledge and understanding of the novel material. This thesis contributes to both
of these areas facilitating further understanding of the possibilities and limitations
associated with the textile architecture as reinforcement in composite materials.
Paper A
An initial experimental study is presented where the mechanical properties of 3Dweave reinforced composites are compared with corresponding properties of 2Dlaminates. The conclusion from Paper A is that the out-of-plane properties are
enhanced, while the in-plane stiffness and strength is reduced. The difference in
strength was not possible to quantify since the specimens with 3D woven material
exhibited other failure modes than those tested for.
Paper B
In Paper B the influential crimp parameter is investigated and three analytical
models are proposed. The warp yarns exhibit 3D crimp, and the paper concludes
that the Youngs modulus decrease non-linearly with increasing crimp. The three
models have different levels of detail, and the more sophisticated models generate
more reliable predictions. However, the overall trends are consistent for all models.
Paper C and D
A novel framework for constitutive modelling of composites reinforced with 3Dwoven preforms is presented in Papers C and D. The framework enables predictive
modelling of both internal architecture and mechanical properties of 3D textile reinforced composites using only a few input parameters. A methodology to generate
near authentic geometry models of 3D-textiles using the concept of inflatable tubes
is presented in Paper C. The result is geometry models which are near authentic
with a high level of detail in features compared with real composite specimens.
The proposed methodology is therefore the main contribution of this thesis to the
field of composite material simulation. In Paper D, the geometries are utilised to
create meso-scale FE-models to calculate the homogenised elastic properties. The
generated FE-models could also be used for progressive damage or permeability
simulations.
Paper E
In the final paper the effect of crimp and different internal weave architectures
is explored. The reinforcement architecture varies between the surfaces and the
interior of the reinforcement, which affects the local stiffness. Both stiffness and
strength decrease non-linearly with increasing crimp. However the strength measure
should be interpreted with care since the stress strain curve of 3D-weave reinforced
specimens exhibit non-linearities believed to be associated with matrix shear cracks.

Introduction

31

Future work

Since the use of 3D-woven pre-forms in textile composites is new, there are many
opportunities for future work. In this thesis the focus has been on the properties
on a meso-scale. It would be interesting to implement the acquired knowledge and
developed tools from this work on full scale beams.
Another area of future work is weave design. As previously discussed, the 3Dweaving process is flexible both in terms of weave types (plain, twill and satin et
cetera) and it is also possible to incorporate more functionality into the beams.
Extra reinforcement (stuffer yarns) may be added where needed and also nonstrucural fibres or cables may be incorporated such as optical fibres or electrical
wiring.
The possible benefits of a fully interlaced yarn structure are not yet fully explored. For instance the increase in out-of-plane strength together with the assumed increase in damage tolerance need to be quantified, and to do so new models
and new experimental approaches are needed.

32

Fredrik Stig

Bibliography
[1]

B. Griffiths. Boeing sets pace for composite usage in large civil aircraft. High
Performance Composites, May 2005.

[2]

P. Apostolopoulos and C. Kassapoglou. Recurring cost minimization of composite laminated structures - optimum part size as a function of learning curve
effects and assembly. Journal of Composite Materials, 36(4):501 518, 2002.

[3]

T. Roberts. The Carbon Fibre industry Worldwide 2011-2020. Material Technology Publications, 2011.

[4]

N.K. Naik and V.K. Ganesh. Prediction of on-axes elastic properties of plain
weave fabric composites. Composites Science and Technology, 45(2):135 152,
1992.

[5]

L. Tong, A.P. Mouritz, and M.K. Bannister. 3D Fibre Reinforced Polymer


Composite. Elsevier, 2002.

[6]

A. K. Prichard. Potential for 3D fabrics in aircraft composites. In Proceedings


of The 3rd World Conference on 3D Fabrics and Their Applications, 2011.

[7]

MOJO. Modular joints for composite aircraft components contract no.:


030871. Annex I - Description of Work.

[8]

J. Ekh. Multi-Fastener Single-lap Joints in Composite Structures. PhD thesis,


Royal Institute of Technology (KTH), 2006.

[9]

M. McCarthy. Bojcas: bolted joints in composite aircraft structures. Air &


Space Europe, 3(3-4):139 142, 2001.

[10] MOJO deliverable D6.2.1 - report of cost and weight assessment. Technical
report, Eurocopter Deutschland, 2009.
[11] P. Linde. Virtual testing of stiffened composite aircraft panels at airbus. In Proceedings of 20th SICOMP Conference on Manufacturing and Design of Composites, Pite, Sweden, 2009.
[12] A. F. Kregers and Yu. G. Melbardis. Plasticity, creep, and rheology of solids.
Mekhanika Polimerov (translated), (1):38, 1978.
[13] C. McHugh. Taking textiles to the next dimension. Aerospace Manufacturing,
July 2011.
[14] N. Khokar. Second-generation woven profiled 3D fabrics from 3D-weaving. In
First World Conference on 3D Fabrics, 2008.

Introduction

33

[15] M.A. Nicosia, F. Vineis, J. Lawrence, and S.T. Holmes. Microstructural modeling of three-dimensional woven fiber composites. In Proceedings of the 9th
International Conference on Textile Composites - TEXCOMP9: Recent Advances in Textile Composites, 2008.
[16] C.H. Chiu and C.C. Cheng. Weaving method of 3D woven preforms for advanced composite materials. Textile Research Journal, 73(1):3741, 2003. Woven preforms;.
[17] H. Gu and Z. Zhili. Tensile behavior of 3D woven composites by using different
fabric structures. Materials and Design, 23(7):671 674, 2002.
[18] D.S. Ivanov, S.V. Lomov, A.E. Bogdanovich, M. Karahan, and I. Verpoest. A
comparative study of tensile properties of non-crimp 3D orthogonal weave and
multi-layer plain weave e-glass composites. part 2: Comprehensive experimental results. Composites-Part A, 40(8):11441157, 2009.
[19] P.J. Callus, A.P. Mouritz, M.K. Bannister, and K.H. Leong. Tensile properties
and failure mechanisms of 3D woven GRP composites. Composites - Part A,
30(11):12771287, 1999.
[20] B.N. Cox and M.S. Dadkhah. Macroscopic elasticity of 3D woven composites.
Journal of Composite Materials, 29(6):785 819, 1995.
[21] A.P. Mouritz and B.N. Cox. A mechanistic interpretation of the comparative
in-plane mechanical properties of 3d woven, stitched and pinned composites.
Composites-Part A, 41(6):709728, 2010.
[22] B.N. Cox, M.S. Dadkhah, and W.L. Morris. On the tensile failure of 3D woven
composites. Composites-Part A, 27(6):447458, 1996.
[23] B.N. Cox, M.S. Dadkhah, W.L. Morris, and J.G. Flintoff. Failure mechanisms
of 3D woven composites in tension, compression, and bending. Acta Metallurgica et Materialia, 42(12):39673984, 1994.
[24] A.E. Bogdanovich. Multi-scale modeling, stress and failure analyses of 3-d
woven composites. Journal of Materials Science, 41(20):65476590, 2006.
[25] J. Brandt, K. Drechsler, and F.-J Arendts. Mechanical performance of composites based on various three-dimensional woven-fibre preforms. Composite
Science and Technology, 56(3):3816, 1996.
[26] M.P. Rao, B.V. Sankar, and G. Subhash. Effect of Z-yarns on the stiffness and strength of three-dimensional woven composites. Composites-Part
B, 40(6):540551, 2009.

34

Fredrik Stig

[27] S. Lomov, A. E. Bogdanovich, D. S. Ivanov, D. Mungalov, M. Karahan, and


I. Verpoest. A comparative study of tensile properties of non-crimp 3D orthogonal weave and multi-layer plain weave e-glass composites. part 1: Materials,
methods and principal results. Composites-Part A, 40(8):11341143, 2009.
[28] P. Tan, L. Tong, and G.P. Steven. Behavior of 3D orthogonal woven CFRP
composites. Part II. FEA and analytical modeling approaches. CompositesPart A, 31(3):273281, 2000.
[29] L. Tong, P. Tan, and G. P. Steven. Effect of yarn waviness on strength of 3D
orthogonal woven CFRP composite materials. Journal of Reinforced Plastics
and Composites, 21(2):153173, 2002.
[30] N. Khokar. 3D-Weaving and noobing: Characterization of interlaced and noninterlaced 3D fabric forming principles. PhD thesis, Chalmers University of
Technology, 1997.
[31] N. Khokar. 3D-weaving: Theory and practice. Journal of the Textile Institute,
92(1):193207, 2001.
[32] N. Khokar. 3D fabric-forming process: Distinguishing between 2D-weaving,
3D-weaving and an unspecified non-interlacing process. Journal of the Textile
Institute, 87(1):97106, 1996.
[33] N. Khokar. Noobing: A nonwoven 3D fabric-forming process explained. Journal of the Textile Institute, 93(1):5274, 2002.
[34] N. Khokar. A classification of shedding methods. Journal of the Textile Institute, 90(4), 1999.
[35] T.J. Whitney and T.W. Chou. Modeling of 3-D angle-interlock textile structural composites. Journal of Composite Materials, 23:890911, 1989.
[36] J. J. Crookston, S. Kari, N. A. Warrior, I. A. Jones, and A. C. Long. 3D textile
composite mechanical properties prediction using automated FEA of the unit
cell. In Proceedings of The Sixteenth International Conference on Composite
Materials, July 8-13, 2007: Kyoto, Japan, 2007.
[37] A. E. Bogdanovich. Advancements in manufacturing and applications of 3-D
woven preforms and composites. In Proceedings of The Sixteenth International
Conference on Composite Materials, July 8-13, 2007: Kyoto, Japan, 2007.
[38] A. Hao, B. Sun, Y. Qiu, and B. Gu. Dynamic properties of 3-D orthogonal woven composite T-beam under transverse impact. Composites-Part A,
39(7):10731082, 2008.
[39] S. Rudov-Clark, A.P. Mouritz, L. Lee, and M.K. Bannister. Fibre damage in the manufacture of advanced three-dimensional woven composites.
Composites-Part A, 34:963970, May 2003.

Introduction

35

[40] M. H. Mohamed, A. E. Bogdanovich, L. C. Dickinson, J. N. Singletary, and


R. Bradley Lienhart. A new generation of 3D woven fabric preform and composite. SAMPE Journal, 37(3):817, 2001.
[41] L.C. Dickinson, M.H. Mohamed, and A.E. Bogdanovich. 3D weaving: What,
how, and where. In Proceedings of the 44th SAMPE Symposium, May 1999.
[42] E. Alexander, M. Lewin, H.V. Muhasm, and M. Shiloh. Definition and measurement of crimp of textile fibres. Textile Research Journal, 26(8):606617,
1956.
[43] F.T. Peirce. Geometry of cloth structure. Journal of the Textile Institute,
28(3):4596, 1937.
[44] A. C. West and D. O. Adams. Axial yarn crimping effects in braided composite
materials. Journal of Composite Materials, 33:402419, 1999.
[45] M. Guagliano and E. Riva. Mechanical behaviour prediction on plane weave
composites. Journal of Strain Analysis for Engineering Design, 36(2):153162,
2001.
[46] T.A. Bogetti, J.W. Gillespie JR, and M.A. Lamontia. Influence of ply waviness
on the stiffness and strength reduction on composite laminates. Journal of
Thermoplastic Composite Materials, 5(4):344369, 1992.
[47] M. R. Garnich and G. Karami. Finite element micromechanics for stiffness and
strength of wavy fiber composites. Journal of Composite Materials, 38(4):273
292, 2004.
[48] D. Zenkert and M. Battley. Foundations of Fibre Composites. KTH Aeronautical and Vehicle Engineering, 1996.
[49] N.V. De Carvalho, S.T. Pinho, and P. Robinson. An experimental study of
failure initiation and propagation in 2d woven composites under compression.
Composites Science and Technology, 71(10):1316 1325, 2011.
[50] B. Budiansky and N.A. Fleck. Compressive failure of fibre composites. Journal
of the Mechanics and Physics of Solids, 41(1):183211, 1993.
[51] S. Lomov, A. E. Bogdanovich, M. Karahan, D. Mungalov, and I. Verpoest.
Internal structure, mechanical properties and damage behaviour of non-crimp
3D orthogonal woven composites. In Proceedings of the 3rd World Conference
on 3D Fabrics and Their Applications, 2011.
[52] S.V. Lomov, A.V. Gusakov, G. Huysmans, A. Prodromou, and I. Verpoest.
Textile geometry preprocessor for meso-mechanical models of woven composites. Composites Science and Technology, 60(11):2083 2095, 2000.

36

Fredrik Stig

[53] A. Adumitroaie and E. J. Barbero. Beyond plain weave fabrics - i. geometrical


model. Composite Structures, 93(5):1424 1432, 2011.
[54] L.Y. Li, P.H. Wen, and M.H. Aliabadi. Meshfree modeling and homogenization of 3d orthogonal woven composites. Composites Science and Technology,
71(15):1777 1788, 2011.
[55] M. Sherburn. Geometric and Mechanical Modelling of Textiles. PhD thesis,
The University of Nottingham, July 2007.
[56] S.V. Lomov, D.S. Ivanov, I. Verpoest, M. Zako, T. Kurashiki, H. Nakai, and
S. Hirosawa. Meso-FE modelling of textile composites: Road map, data flow
and algorithms. Composites Science and Technology, 67(9):18701891, 2007.
[57] I. Verpoest and S. V. Lomov. Virtual textile composites software WiseTex: Integration with micro-mechanical, permeability and structural analysis. Composites Science and Technology, 65(15-16 SPEC ISS):25632574, 2005.
[58] G. Hivet and P. Boisse. Consistent 3D geometrical model of fabric elementary
cell. application to a meshing preprocessor for 3D finite element analysis. Finite
elements in analysis and design, 42:2549, 2005.
[59] F. Robitaille, A. C. Long, I. A. Jones, and C. D. Rudd. Automatically generated geometric descriptions of textile and composite unit cells. CompositesPart A, 34(4):303312, 2003.
[60] S. Z. Sheng and S.V. Hoa. Modeling of 3D angle interlock woven fabric composites. Journal of Thermoplastic Composite Materials, 16(1):4558, 2003.
[61] Z. Wu. Three-dimensional exact modeling of geometric and mechanical properties of woven composites. Acta Mechanica Solida Sinica, 22(5):479486, 2009.
[62] P. Tan, L. Tong, and G.P. Steven. Modeling approaches for 3D orthogonal
woven composites. Journal of Reinforced Plastics and Composites, 17(6):545
577, 1998.
[63] W.R. Walsh L.P. Djukic, I.Herszberg and G.A. Schoeppner. Fabric and
resin additives for contrast enhancement in visualisation of woven composite tow architecture using a microCT scanner. Composites-Part A,
10.1016/j.compositesa.2008.12.016, 2009.
[64] F. Desplentere, S.V. Lomov, D.L. Woerdeman, I. Verpoest, M. Wevers, and
A. Bogdanovich. Micro-CT characterization of variability in 3D textile architecture. Composites Science and Technology 65 (2005) 19201930, 65:1920
1930, 2005.
[65] M. Sherburn, A. Long, and A. Jones. Prediction of textile geometry using
strain energy minimisation. In Proceedings of the 1st World Conference On
3D Fabrics, 2008.

Introduction

37

[66] T.V. Sagar, P. Potluri, and J.W.S. Hearle. Mesoscale modelling of interlaced fibre assemblies using energy method. Computational Material Science,
28(1):4962, 2003.
[67] Y. Mahadik and S.R. Hallett. Finite element modelling of tow geometry in 3D
woven fabrics. Composites-Part A, 41(9):11921200, 2010.
[68] D. Durville. Simulation of the mechanical behaviour of woven fabrics at the
scale of fibers. International Journal of Material Forming, 2010.
[69] R. M. Christensen. Mechanics of Composite Materials. Dover Publications
Inc, 2005.
[70] T. Mori and K. Tanaka. Average stress in matrix and average elastic energy of
materials with misfitting inclusions. Acta Metallurgica, 21(5):571 574, 1973.
[71] B. Gommers, I. Verpoest, and P. van Houtte. Mori-Tanaka method applied to
textile composite materials. Acta Materialia, 46(6):2223 2235, 1998.
[72] Ph. Vandeurzen, J. Ivens, and I. Verpoest. Micro-stress analysis of woven fabric composites by multilevel decomposition. Journal of Composite Materials,
32(7):623 651, 1998.
[73] A. G. Prodromou, S. V. Lomov, and I. Verpoest. The method of cells and the
mechanical properties of textile composites. Composite Structures, 93(4):1290
1299, 2011.

38

Fredrik Stig

Division of work between authors


Paper A
Stig performed the weaving and testing and wrote the paper. Hallstrm initiated
and guided the work and contributed to the paper with valuable comments and
revisions.
Paper B
Stig developed the analytical models and wrote the paper. Hallstrm initiated
and guided the work and contributed to the paper with valuable comments and
revisions.
Paper C
Stig developed the geometry modelling principles, implemented analysis scheme
and wrote the paper. Hallstrm initiated and guided the work and contributed to
the paper with valuable comments and revisions.
Paper D
Stig developed the modelling frame work and wrote the paper. Hallstrm initiated
and guided the work and contributed to the paper with valuable comments and
revisions.
Paper E
Stig performed the experiments, analysed the data and wrote the paper. Hallstrm
initiated and guided the work and contributed to the paper with valuable comments
and revisions.

Potrebbero piacerti anche