Vous êtes sur la page 1sur 257

QUASI-STATIC AND IMPACT

CHARACTERIZATIONS OF STITCHED FLAX

FIBER COMPOSITES

MOHAMMAD RAVANDI

(M.Sc. (Hons), Amirkabir University of Technology, Iran)

A THESIS SUBMITTED
FOR THE DEGREE OF DOCTOR OF PHILOSOPHY
DEPARTEMENT OF MECHANICAL ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE

2017

Supervisors:
Associate Professor Yong Ming Shyan John, Main Supervisor
Professor Tay Tong Earn, Co-Supervisor
Dr. Teo Wern Sze, Co-Supervisor
Examiners:
Associate Professor Tan Beng Chye, Vincent
Associate Professor Lim Kian Meng
Professor Ignace Verpoest, University of Leuven
DECLARATION

I hereby declare that this thesis is my original work and it has been written by

me in its entirety. I have duly acknowledged all the sources of information

which have been used in the thesis.

This thesis has also not been submitted for any degree in any university

previously.

Mohammad Ravandi

20 April 2017

i
Acknowledgements

Foremost, I would like to express my sincere gratitude and appreciation to my

supervisors Prof. Tay Tong Earn, Dr. Yong Ming Shyan and Dr. Teo Wern Sze

for their continuous support, motivation, guidance and advice throughout my

PhD journey. It would never have been possible for me to have this thesis

accomplished without all of their invaluable supports. Besides my supervisors,

I would like to thank Dr. Tran Le Quan Ngoc for sharing his knowledge and

experiences on natural fiber composites as well as his helpful discussions.

The list of people I have to thank is really endless, but I could not have it done

without a single one of them. I express my warm thanks to Dr. Ridha

Muhammad for his valuable help and guidance on the finite element modeling,

Dr. Deng Xinying for her countless help with carrying out experiments at

SIMTech, and Dr. Umeyr Kureemun for his useful discussions on the testing

and modeling of composites. Many thanks to lab technologists Mr. Abdul Malik

Bin Baba, Mr. Low Chee Wah and Mr. Chiam Tow Jong for their all great

assistance with lab equipment, and providing precious technical supports during

the long hours of lab work.

My heartfelt thanks to all my friends in applied mechanics group of National

University of Singapore, Dr. Boyang Chen, Dr. Abhishek Vishwanat, Dr. Andy

Haris, Mr. Ashish Kumar, Mr. M.R. Abir , Dr. Hu Xiaofei, Dr. Nader Hamzavi,

Dr. Saeid Arabnejad, Mr. Ebrahim Sadeghpour, and all my fellow labmates in

ii
SIMTech, Dr. Steve Phillips, Dr. Zhang Xiwen, Ms Tan Cher Lin, for your

kindness, friendship and unwavering support that you have brought me.

Special thanks to the Singapore international graduate award and A*STAR for

funding this project and the four-year scholarship.

Last but not the least, I wish to express my deepest gratitude to my parents,

Hajar Ravandi and Maghsoud Ravandi for your endless love, support,

encouragement and trust in my life. And to Nastaran Aslnajjari, thank you for

being in my life, for your love and company, for your understanding and support,

as well as for all of the sacrifices you made for me.

iii
Table of Contents

Chapter 1 Introduction ................................................................................ 1


1.1. Background ......................................................................................... 1
1.2. Problem statement ............................................................................... 3
1.3. Objectives of the research ................................................................... 4
1.4. Structure of the thesis .......................................................................... 5

Chapter 2 A Literature review.................................................................... 8


2.1. Introduction ......................................................................................... 8
2.2. Trend in using plant fibers to reinforce polymers ............................... 9
2.3. Flax fiber composites ........................................................................ 13
2.3.1. Flax fibers .................................................................................. 13
2.3.2. Flax fiber reinforced composites ............................................... 17
2.3.3. Fiber Architecture ...................................................................... 19
2.3.4. Thermoset resins ........................................................................ 24
2.3.5. Fiber-matrix bonding ................................................................. 25
2.3.6. Composite processing ................................................................ 26
2.4. Impact behavior of laminated composites ......................................... 28
2.4.1. Composite laminates under low-velocity impact ....................... 28
2.4.2. Low-velocity impact behavior of natural fiber composites ....... 32
2.5. Through-the-thickness stitched composites ...................................... 35
2.5.1. Effect of stitching on mechanical properties and toughness of
composites.................................................................................. 38
2.5.2. Effect of stitching on impact behavior of composites ............... 45
2.6. Numerical analysis of stitched and woven composites ..................... 47
2.6.1. Numerical modeling for natural fiber composites ..................... 48
2.6.2. Predicting interlaminar fracture toughness of stitched
composites................................................................................. 50
2.6.3. Failure analysis of woven composites under impact loads ........ 55
2.7. Summary ........................................................................................... 57

Chapter 3 Experimental Methodology ..................................................... 59


3.1. Introduction ....................................................................................... 59
3.2. Material ............................................................................................. 60

iv
3.2.1. In-plane flax fiber reinforcements ............................................. 60
3.2.2. Stitch fibers ................................................................................ 60
3.2.3. Thermoset matrix ....................................................................... 61
3.3. Preform stitching ............................................................................... 62
3.3.1. Cotton thread stitching ............................................................... 62
3.3.2. Flax yarn stitching...................................................................... 63
3.4. Composite manufacturing ................................................................. 66
3.5. Fiber volume fraction measurement .................................................. 69
3.6. Composite tensile testing .................................................................. 69
3.7. Strength evaluation of the stitch fibers .............................................. 70
3.8. Flexural testing .................................................................................. 71
3.9. Composite double cantilever beam (DCB) testing ............................ 72
3.10. Low-velocity impact testing .......................................................... 75
3.11. Scanning Electron Microscopy ...................................................... 77

Chapter 4 Mechanical and fracture properties of stitched


flax fiber composites ................................................................ 79
4.1. Introduction ....................................................................................... 79
4.2. Strength of dry and impregnated stitch fibers ................................... 80
4.3. Effects of stitch areal fraction on the in-plane properties
of the woven composite..................................................................... 82
4.3.1. Tensile properties ....................................................................... 84
4.3.2. Flexural properties ..................................................................... 88
4.4. Effect of stitch areal fraction on the Mode I interlaminar
fracture toughness of flax woven fabric composite .......................... 91
4.5. Fracture mechanisms of Mode I DCB samples from SEM
micrographs ....................................................................................... 96
4.6. Effect of in-plane fiber architecture on the Mode I interlaminar
fracture toughness of unstitched and stitched
composite laminates .......................................................................... 98
4.7. Conclusion ....................................................................................... 103

Chapter 5 Numerical modeling of Mode I interlaminar failure of


stitched flax fiber composites ................................................ 105
5.1. Introduction ..................................................................................... 105
5.2. DCB test simulation of stitched flax fiber composite ..................... 106
5.2.1. Modeling strategy .................................................................... 106
5.2.2. Material Model......................................................................... 110

v
5.3. Modeling of in-plane fiber bridging by means of cohesive law ..... 116
5.3.1. Testing for fracture properties of unstitched composite .......... 116
5.3.2. Superposition of cohesive laws for modeling the effect
of fiber bridging ....................................................................... 119
5.3.3. FEM modeling of DCB test of unstitched composite .............. 123
5.4. Results of FEM modeling of DCB testing of stitched composite ... 127
5.4.1. Parametric study....................................................................... 134
5.5. Conclusion ....................................................................................... 136

Chapter 6 Low-velocity impact performance of stitched flax/epoxy


composite laminates ............................................................... 138
6.1. Introduction ..................................................................................... 138
6.2. Specimen types and the drop-weight impact test parameters ......... 139
6.3. Visual inspection of post-impact specimens ................................... 141
6.3.1. Non-perforation impact test (19.6J) ......................................... 141
6.3.2. Perforation impact test (44.6J) ................................................. 143
6.4. Impact force- time history ............................................................... 144
6.5. Impact force-displacement response ............................................... 151
6.6. Energy absorption and impact fracture toughness .......................... 155
6.7. Conclusion ....................................................................................... 162

Chapter 7 Numerical analysis of low-velocity impact on woven


flax/epoxy composite laminate .............................................. 164
7.1. Introduction ..................................................................................... 164
7.2. Modeling strategy ............................................................................ 165
7.3. Material model ................................................................................ 168
7.3.1. Elastic behavior of woven laminate ......................................... 169
7.3.2. Fiber failure .............................................................................. 171
7.3.3. Matrix failure ........................................................................... 172
7.3.4. Delamination ............................................................................ 174
7.4. Numerical simulation results and discussion .................................. 175
7.4.1. Impact force response .............................................................. 175
7.4.2. Impact damage evolution ......................................................... 178
7.4.3. Matrix damage and delamination............................................. 184
7.4.4. Permanent deformation ............................................................ 185
7.4.5. Effect of intralaminar fracture toughness ................................ 187
7.5. Conclusion ....................................................................................... 188

vi
Chapter 8 General Conclusion and recommendations ......................... 190
8.1. General Conclusion ......................................................................... 190
8.2. Recommendations ........................................................................... 195

References ……………………………………………………………………….. 199


Appendix A ………………………………………………………...…………….. 220
Appendix B ...…………………………………………………………………….. 225
Appendix C ...…………………………………………………………………….. 228

vii
Summary

Flax fibers, desired for their high specific stiffness, vibration damping, and low

environmental impact, make suitable reinforcements in polymer composites for

moderate load-bearing applications. The composite mechanical properties are

able to match those of short random glass fiber reinforced polymers when the

flax fibers are used in the form of laminates of continuous uni-directional (UD)

or woven fibers made from optimally twisted yarns. To improve the out-of-

plane properties and damage tolerance of conventional laminated composites,

stitching has been recognized as a convenient and cost-effective technique.

However, the effectiveness of stitching natural fiber preforms to improve the

performance of resin infused natural fiber laminated composites have yet to be

widely investigated. Modeling of damage development in continuous natural

fiber composites is also fairly uncommon at this moment.

This thesis reports a comprehensive study on the effects of through-the-

thickness stitching using natural fibers on the mechanical and impact properties

of flax fiber/epoxy composite laminates. Woven fabric and unidirectional

laminates preforms stitched with twistless flax yarn and twisted cotton threads

were studied. The tensile, flexural and Mode I interlaminar fracture toughness

of composites with various stitch areal densities were experimentally

determined, followed by transverse impact tests on the optimally stitched woven

fiber composites. A finite element model of a double-cantilever-beam test was

successfully developed to give accurate predictions of delamination propagation

in stitched natural fiber composites, taking into account effects of the extensive

viii
in-plane fiber bridging and stitch yarn by means of trilinear cohesive law and 2-

node beam elements, respectively. A finite element modeling of the drop-weight

impact test on woven flax fiber composite was also implemented to study the

material behavior, damage development, and impact fracture resistance.

The results of this study show that flax-yarn stitches perform far better than

cotton threads to enhance the quasi-static interlaminar fracture toughness of

flax/epoxy composites. However, the stitch-induced defects also lowered the

tensile properties and transverse impact resistance of the composites

substantially. Damage development tends to be dominated by both fiber and

matrix fractures, resulting in the fracture and energy absorption mechanisms

being quite different from synthetic fiber composites. Finally, the numerical

work showed that the continuum damage mechanics model is able to predict the

force history, deformation and the impact damage of woven flax composite with

good accuracy.

ix
List of Symbols

𝐴𝑠 Average cross-section area of the stitch


𝑎 Delamination crack length
∆𝑎 Change in crack length
𝑏 Specimen width
𝐶, 𝐾 Parameters of hardening function
𝐷𝑖𝑗 Penalty stiffness of cohesive element
𝑑𝑖𝑗 Stiffness degradation factor in ij-direction
𝐸𝑖𝑚𝑝𝑎𝑐𝑡 Impact energy
𝐸𝑎𝑏𝑠𝑜𝑟𝑏𝑒𝑑 Absorbed energy of impact
E11 Longitudinal Young’s modulus
E22 Transverse Young’s modulus
E33 Out-of-plane Young’s modulus
𝐹𝑖𝑗 Damage activation function in ij-direction
G12 , G13 In-plane shear modulus
G23 Out-of-plane shear modulus
𝐺𝐶𝑚 Mixed-mode fracture energy
𝐺𝐼𝑐 Mode I fracture toughness
𝐺𝐼𝐼𝑐 Mode II fracture toughness
𝐺1 Initiation value of fracture toughness
𝐺2 Subtraction of propagation and initiation values of 𝐺𝐼𝐶
𝐺𝑐 Propagation values of fracture toughness
𝐺𝑓+ Tensile fracture toughness of woven composites in fiber
directions
𝐺𝑓−𝛼 Compressive fracture toughness of woven composites in fiber
directions
𝐺𝑖𝑚𝑝𝑎𝑐𝑡 Impact fracture toughness
𝑔 Acceleration of gravity
𝐿 Length of through-thickness crack in laminates
𝑙𝑐 Material characteristic length
𝑙𝑐_𝑆𝑢𝑝 Characteristic length of the process zone for the trilinear
cohesive law
𝑠𝑠
𝑙𝑝𝑧_𝑆𝑢𝑝 Steady-state process zone lengths for the trilinear cohesive law
𝑀 Mass of impactor
𝑚, 𝑛 Superposition parameters of trilinear cohesive law
𝑃 Crack opening load
𝑃1 Damage initiation force under impact loading
𝑃𝑚𝑎𝑥 Maximum impact force

x
𝑃𝑖𝑚𝑝 Impact force
𝑠𝑡𝑖𝑡𝑐ℎ Stitch cross-sectional radius
𝑅
𝑟𝑖𝑗 Damage thresholds in ij-direction
𝑆 Shear yielding stress
𝑆𝑅 Stitch row spacing
𝑆𝐿 Stitch length
𝑡𝑝 Thickness of a single ply
𝑡𝑎𝑟𝑚 Thickness of plies stacking with same orientation
𝑉𝑐 Contact velocity in impact
𝑉𝑓 Fiber volume fraction
𝑋𝑓+ Ultimate tensile strength of woven composites in fiber directions
𝑋𝑓− Ultimate compressive strength of woven composites in fiber
directions
𝑋𝑡𝑠𝑡𝑖𝑡𝑐ℎ Strength of stitch yarn
𝛿𝑖𝑚𝑝 Impactor displacement
∆ Crack length correction factor
𝛿 Crack opening displacement
𝛿eff Effective displacement
𝜖𝑖𝑗 Strain components
𝑒𝑙 In-plane elastic shear strain
𝜀12
𝑝𝑙 In-plane plastic shear strain
𝜀12
𝜎̃ Effective stress
𝜎𝑖𝑗 Stress components
𝜎𝑠 , 𝜎𝑡 Shear stress components
𝜎𝑛𝑐 Normal strength of cohesive interface
𝜎𝑠𝑐 , 𝜎𝑡𝑐 Shear strength of cohesive interface
𝜎̃12 Effective shear stress
𝜙𝑖𝑗 Ratio of stress 𝜎𝑖𝑗 to strength 𝑋𝑖𝑗
𝜂 Material parameter in the B-K criterion
𝑣 Poisson’s ratio
𝜌𝑚 Density of pure cured matrix
𝜌𝑓 Density of fiber
𝜌𝑐 Density of composite

xi
List of Abbreviations

1D One-dimensional
2D Two-dimensional
3D Three-dimensional
ASTM American society for testing and materials
B31 Abaqus 2-node beam element
CAI Compression after impact
CC Compliance calibration
CDM Continuum damage mechanics
CFRP Carbon fiber reinforced plastics
COD Crack opening displacement
COH3D8 Abaqus 8-node 3D cohesive element
DCB Double cantilever beam
ENF End notched flexure
FAI Flexure after impact
FE Finite element
FEM Finite element method
FFT Fast Fourier transform
FRP Fiber reinforced plastics
FT Fracture toughness
MBT Modified beam theory
MCC Modified compliance calibration
MFA Microfibrillar angle
NL Deviation from linearity in load-displacement curve of DCB test
PTFE Polytetrafluoroethylene
RTM Resin transfer molding
RVE Representative volume element
SEM Scanning electron microscopic
SC8R Abaqus 8-node quadrilateral continuum shell element
UD Unidirectional
UMAT User-defined material model in Abaqus/Standard
VARI Vacuum Assisted Resin Infusion
VARTM Vacuum assisted resin transfer molding
VCCT Virtual crack closure technique
VIS Initiation of fracture obtained by visual observation during DCB test
VUMAT User-defined material model in Abaqus/Explicit

xii
List of Tables

Table 2.1. General advantages of natural fibers over synthetic glass fiber [37].

Table 2.2. Chemical, morphological and mechanical properties of some

natural fibers used for reinforcing polymers [1,39,45,57–62].

Table 3.1. Summary of the composite lay-up and stitch parameters of the

manufactured composite laminates.

Table 4.1. Tensile properties of cotton thread and flax twist-less yarn.

Table 5.1. Material properties used in the numerical model of DCB test.

Table 6.1. The composite lay-up, notation and stitch parameters used for the

manufactured flax fiber laminates for the drop weight impact test.

Table 6.2. Information of Impact test and specimen’s thickness.

Table 6.3. Summary of impact test results.

Table 7.1. Material properties used in numerical modeling.

xiii
List of Figures

Figure 1.1. Schematic outline of this thesis.

Figure 2.1. Classification of reinforcing fibers (reproduced from [40]).

Figure 2.2. The performance and corresponding applications of composites

made of natural fiber (flax) in different fiber forms (reproduced form [44]).

Figure 2.3. Schematic and SEM images of technical fibers versus elementary

fibers (adapted from [44]).

Figure 2.4. Hierarchal structure of flax fiber from stem to microfibril (figures

were reproduced from [44,54]).

Figure 2.5. Ashby plots comparing (a) absolute tensile properties, (b) tensile

properties per unit density for different natural fibers and E-glass [5].

Figure 2.6. Comparison of the absolute and specific tensile properties of

natural fiber composites with E-glass composites [5].

Figure 2.7. Different flax fiber processing and selection of fiber preforms for

composite [47].

Figure 2.8. (a) Effect of twist on yarn tensile strength and failure mechanism;

(b) Tensile strength of (●) long and (□) short flax fiber epoxy impregnated

yarns as a function of twist level [74].

Figure 2.9. Comparison of static and post-fatigue tensile properties of flax–

epoxy composite after exposing to fatigue cycling up to 500,000 cycles, (a) the

tensile modulus and (b) the ultimate tensile strength [8].

Figure 2.10. Classification of natural fiber treatment methods [1].

xiv
Figure 2.11. The performance of various natural fiber composites

manufactured using different techniques [46].

Figure 2.12. Typical impact load–time and impact energy–time histories of (a)

non-penetrated and (b) penetrated laminate [87].

Figure 2.13. Schematic illustration of a typical impact damage mode for

composite laminate [87].

Figure 2.14. Example of (a) complex preform manufacturing through

stitching and (b) commercially available robotic stitching machine (Altin

Nahtechnik GmbH) [21].

Figure 2.15. Examples of stitching damages; (a) in-plane fiber breakage, (b)

local misalignment of fibers around a stitch and (c) schematic of fiber

crimping [21,108].

Figure 2.16. Diagrammatic side view of the sewing machine and its particular

needle developed for the brittle thread (reproduced from [112]).

Figure 2.17. Schematic illustration of (a) lock stitch, (b) modified lock stitch

and (c) chain stitch.

Figure 2.18. DCB test of (a) unstitched and (b) stitched specimen; (c) Load

versus COD curves of unstitched and stitched composites with similar fiber

orientation [108].

Figure 2.19. Stitch fracture process: (a) Initial stitch; (b) Interfacial debonding;

(c) Slack absorption; (d) Stitch fracture and (e) Frictional pull-out [27].

Figure 2.20. Effect of increasing stitch density on the tensile strength and

modulus of chain stitched (a) S2 glass/polyester and (b) Kevlar/PVB-Phenol

composites [105].

xv
Figure 2.21. Effect of z-binder content (stitch density) on the mechanical

properties of various synthetic fiber composites; (a) normalized tensile

strength, (b) normalized tensile modulus, (c) normalized flexural strength, (d)

normalized flexural modulus (The normalized properties, defined as the ratio

of properties of stitched composites to that of the equivalent unstitched

laminate) [32].

Figure 2.22. (a) Schematic of the unit cell used in [139]; (b) Microstructures of

short flax fibers in matrix with different orientation coefficient [140].

Figure 2.23. Illustration of an embedded stitch fiber (a) during elastic

stretching and (b) after slippage of embedded end without interconnected

stitched effect; (c) interconnected stitched fiber during elastic stretching [110].

Figure 2.24. (a) Interlaminar tension test specimen proposed by [27]; (b)

fractured stitch after interlaminar tension test; (c) fractured stitches at the

surface of the test specimen after DCB test [23].

Figure 2.25. Modeling of z-fiber mechanical properties [23].

Figure 3.1. (a) Biotex flax 4x4 hopsack fabric; (b) FlaxTape©110

unidirectional fiber.

Figure 3.2. (a) Biotex 250tex flax yarn; (b) Natural cotton C Ne 50 thread.

Figure 3.3. (a) Stitching using sewing machine; (b) Cross-section of the

cotton-thread stitched woven flax/epoxy composite.

Figure 3.4. Cross-section of the flax-yarn stitched woven flax/epoxy

composite.

Figure 3.5. (a) Schematic view of a stitched preform of woven flax fiber; (b)

Definition of stitch parameters shown from side and top view.

xvi
Figure 3.6. The vacuum assistant resin infusion setup used for manufacturing

composite laminates.

Figure 3.7. (a) and (b) Optical micrograph of the cross-section of the flax fiber

composite; (c) SEM image of a fractured resin-impregnated flax fiber in the

composite.

Figure 3.8. Tensile test of the woven flax fiber composite.

Figure 3.9. (a) Schematic view of the mounting tab and (b) the tensile test of

stitch fiber specimen.

Figure 3.10. Three-point bending test of the woven flax fiber composite.

Figure 3.11. Schematic of tabbed double cantilever beam (DCB) test

specimen.

Figure 3.12. DCB test setup and mounted specimen.

Figure 3.13. Drop-weight impact test setup with a mounted woven flax

specimen.

Figure 4.1. (a) Comparison of load-elongation curves of cotton thread and flax

yarn before and after resin impregnation; (b) dry twist-less flax yarn with

binding thread; (c) dry twisted cotton thread.

Figure 4.2. Effects of various stitch areal fractions of cotton thread and flax

yarn on the tensile properties of woven composite; (a) tensile strength; (b)

tensile stiffness.

Figure 4.3. Effects of various stitch areal fractions of cotton thread and flax

yarn on the flexural properties of the woven composite; (a) flexural strength;

(b) flexural stiffness.

Figure 4.4. Schematic of three-point bending test of woven flax composite.

xvii
Figure 4.5. Failure of (a) Flax yarn and (b) Cotton thread stitched woven flax

composite under bending stress.

Figure 4.6. Effects of stitch areal fraction and stitch material on the

interlaminar fracture toughness (GIC) of woven flax/epoxy composites.

Figure 4.7. R-curves of the stitched flax woven laminates using cotton thread

and flax yarn at almost equal stitch areal fraction of 0.02.

Figure 4.8. Comparison of the frequency of the stitch-induced defects at a

similar stitch areal fraction of (a) cotton thread and (b) flax yarn stitched

composites.

Figure 4.9. SEM image of fractured stitches within the stitched woven

composite; (a) cotton thread, (b) flax yarn.

Figure 4.10. The 3-step delamination growth in the stitched flax fiber

composites.

Figure 4.11. Typical Mode I load-displacement curves of unstitched and

stitched flax UD and woven laminates.

Figure 4.12. Comparison of R-curves of the flax and glass fiber composites.

Figure 4.13. Fiber bridging during delamination growth; (a) woven flax fabric,

(b) unidirectional flax, and (c) unidirectional glass fiber laminate composite.

Figure 4.14. SEM of a fractured surface of the flax woven fabric composite

after DCB test; (a) fiber pull out and breakage, (b) Fiber-matrix debonding,(c)

a resin-rich region between warp and weft fibers, (d) matrix shear yielding.

Figure 5.1. Schematic of DCB specimen of the stitched UD flax fiber

composite, and stitching parameters.

Figure 5.2. 3D finite element model of the DCB test specimen. A finer mesh

was used at the stitched region.

xviii
Figure 5.3. Three fracture mechanisms involved in delamination propagation.

Figure 5.4. Superposition of two bilinear cohesive law and resultant trilinear

cohesive law (reproduce from [151]).

Figure 5.5. SEM of the delaminated surface for the DCB test specimen of (a)

UD and (b) woven flax fiber composite. Fracture of stitch (flax) yarn occurs at

the delamination surface with minimum fiber pull-out.

Figure 5.6. Experimental GIC values of an unstitched UD flax/epoxy

composite calculated using the MBT, CC and MCC methods proposed under

ASTM D5528 [31].

Figure 5.7. In-plane fiber bridging across the delamination plane of UD flax

fiber composite during DCB test.

Figure 5.8. Force–displacement response of DCB test specimen of unstitched

UD flax fiber composite; FEM & Experimental results.

Figure 5.9. Experimental R-curve and FEM analysis results of bilinear and

trilinear cohesive law.

Figure 5.10. Force–displacement response of DCB test specimen of stitched

UD flax fiber composite; Experimental versus FEM analysis results.

Figure 5.11. Experimental and FEM analysis R-curve of stitched UD flax fiber

composite.

Figure 5.12. Change in crack length against cross-head opening displacement

for experimental and FEM analysis

Figure 5.13. The 3-step delamination growth marking on the load-

displacement curve of DCB test of stitched composite; comparison of

prediction with experimental results.

xix
Figure 5.14. Interlaminar damage (delamination) evolution within the stitched

UD flax fiber composite; (a) the cohesive interface failure propagates and

passes round the stitch elements, (b) the crack then is arrested behind the stitch

elements while the stitch elements are stretched, (c) the stitch elements

fractures and the stored energy is released to propagate the cohesive crack at

higher speed until the next row of stitch elements.

Figure 5.15. Effect of variation of stitch strength; (a) Force–displacement

response and (b) R-curve of the stitched UD flax fiber composite.

Figure 6.1. Damage of 19.6J impact (non-perforation) on both faces of (a)

unstitched woven – PW0, (b) cotton stitched woven – PWC and (c) flax

stitched woven – PWF, (d) unstitched cross-ply – UD0.

Figure 6.2. Damage of 44.6J impact (non-perforation) on both faces of (a)

unstitched woven – PW0, (b) cotton stitched woven – PWC and (c) flax

stitched woven – PWF, (d) unstitched cross-ply – UD0.

Figure 6.3. Impact force-time histories of 19.6 J (non-perforation) for (a)

unstitched woven – PW0, (b) unstitched cross-ply – UD0, (c) cotton stitched

woven – PWC and (d) flax stitched woven – PWF.

Figure 6.4. Impact force-time histories of 44.6 J (perforation) for (a)

unstitched woven – PW0, (b) unstitched cross-ply – UD0, (c) cotton stitched

woven – PWC and (d) flax stitched woven – PWF.

Figure 6.5. Typical representative curve of the force-time responses for (a)

woven flax laminates, (b) cross-ply flax laminate.

Figure 6.6. Through-thickness inspection of the impacted woven laminates by

the cross-sectional microscopy.

xx
Figure 6.7. Schematic view of a woven flax composite under impact loading in

which the failure in the fiber layers due to tensile (𝜎𝑇) or compressive (𝜎𝐶)

stress happens prior to a failure in resin-rich inter-ply layers due to shear stress

(τ).

Figure 6.8. Impact force-displacement response of 19.6 J for (a) unstitched

woven – PW0, (b) unstitched cross-ply – UD0, (c) cotton stitched woven –

PWC and (d) flax stitched woven – PWF.

Figure 6.9. Impact force-displacement response of 44.6 J for (a) unstitched

woven – PW0, (b) unstitched cross-ply – UD0, (c) cotton stitched woven –

PWC and (d) flax stitched woven – PWF.

Figure 6.10. Different zones of impact force-displacement response for all the

flax laminates subjected to (a) 19.6J and (b) 44.6J impact energy.

Figure 6.11. Energy absorption curves under 19.6J and 44.6J impact energies.

Figure 6.12. Scanning electron microscopic (SEM) of fracture surfaces of the

woven flax fiber laminate under transverse impact; (a) warp fiber fracture and

weft fiber/matrix debonding at the fracture surface, (b) fiber bundle fracture,

(c) and (d) river marking, cleavage marking and cracking of matrix.

Figure 6.13. Effect of stitch fiber on damage initiation load and fracture

toughness under impact load.

Figure 7.1. Schematic of (a) the model dimension, and (b) the lay-up used.

Figure 7.2. 3D finite element model of the drop-weight impact test.

Figure 7.3. Schematic of 2D woven fabric reinforcement and the local

coordinate system.

xxi
Figure 7.4. Monotonic and cyclic shear stress-strain curves obtained from

tensile tests on ±45 woven flax/epoxy laminate.

Figure 7.5. Impact force–time curves of numerical simulation versus

experimental results for (a) 19.6J (non-perforation) and (b) 44.6J (perforation).

Figure 7.6. Impact force–displacement curves of numerical simulation versus

experimental results for (a) 19.6J (non-perforation) and (b) 44.6J (perforation).

Figure 7.7. Damage evolution at top and back surface of composite plate and

the all cohesive interfaces during low-velocity impact at specified time T1 to

T4 for (a) 19.6 J and (b) 44.6J impact energy (The damage at T4 is compared

with the actual specimen).

Figure 7.8. Comparison of energy dissipation by delamination with matrix

damage in woven flax fiber composite subjected to (a) 19.6J and (b) 44.6J

impact energy.

Figure 7.9. Permanent indentation after non-perforation impact (19.6J); (a)

average experimental measurements = 3.95 mm; (b) numerical results = 3.42

mm.

Figure 7.10. Deformed shape after full perforation impact (44.6J)

Figure 7.11. (a) Impact force history of unstitched and flax yarn stitched

laminates; (b) Prediction of damage on the impact and back surface of the flax

yarn stitched laminate using reduced FT values.

xxii
Chapter 1 1

Chapter 1
Introduction

1.1. Background
In recent decades, fiber reinforced polymer composites (such as carbon and

glass fiber composites) have been utilized widely in various applications in the

aerospace, automotive, marine and construction sectors thanks to their potential

to combine high mechanical properties with low weight. Despite their many

advantages, there are some serious concerns that challenge the continued use of

these conventional composites. They are mostly synthetic, petroleum based and

require large amounts of energy to be produced. The increase in greenhouse gas

emissions and land pollution stemming from disposed polymer and fiber

reinforced polymer composite products have caused serious environmental

problems and climate change. Although it has been agreed between global

governments to manage and reduce the environment impact drastically, it does

not alter the fact that non-renewable fossil fuel from which our energy and

materials derive from will eventually be depleted. Therefore there is a pressing

need to develop sustainable alternative materials that are not only derived from

renewable resources, but also use less energy to produce, have higher structural

performance and have less detrimental impact on the environment.


2 Introduction

Recently, the use of natural plant fiber reinforcements in polymer composite

materials has gained traction due to their promising properties such as high

specific stiffness, good acoustic insulation and vibration damping, and lower

environmental impacts. These advantages of natural fibers make them suitable

for replacing synthetic glass fiber reinforced composites in many nonstructural

and semi-structural applications, and such substitutions based on short, non-

woven natural fiber preforms are becoming increasingly common in the

transportation and sport goods industries [1–3]. There is great interest to extend

the use of natural fiber composites to structural applications, and an increasing

number of investigations are now beginning to research into continuous long

natural fiber fabric systems to elevate the performance of existing composites

[1,4]. For the same reasons, the studies in this thesis have used long, continuous

flax fibers for the in-plane reinforcement.

Among the natural fibers, flax1 has great potential in reinforcing polymers for

structural applications due to its better consistency and higher mechanical

properties [2,5]. Flax fibers, in the form of short fibers or non-woven mats, are

very common and have been widely used in the composite industry for

producing nonstructural components. However, in order to maximize the

strength and stiffness performance, it is necessary to use long continuous

unidirectional (UD) or woven fibers with optimally twisted yarns [6]. In this

regard, the performance of composite laminates made out of long flax fiber

reinforcements, have been the subject of several experimental studies [7–12]

1 Also known with binomial name “Linum usitatissimum”


Chapter 1 3

and the effects of various influencing factors like fiber preform architecture,

stacking sequence, etc., have been investigated and were found to be significant.

1.2. Problem statement


An important failure mode in these 2D-reinforced orthotropic composite

laminates is the interlaminar failure (i.e. delamination) due to the lack of

reinforcing fibers acting through the thickness of the laminates. The weakness

in the interlaminar properties and consequently, the poor transverse

performance of composite laminates, is usually of concern as it is well

established that sub-surface micro-fractures and delamination damage can arise

from relatively low velocity incidents such as tool drops, stone or debris strike,

etc. For this reason, low-velocity transverse impact behavior of composite

laminates has always been one of the main design considerations in many

composite applications [13,14]. When lower-strength reinforcements, like

natural fibers, are being used, this behavior can be more critical due to

contribution of the fiber failure [15], and this is our main motivation for

studying the impact properties and behavior of the flax/epoxy system as well.

Various techniques have been introduced to enhance the interlaminar strength

of composites; for example engineering the interface adhesion between fiber

and matrix, increasing toughness of matrix, interleaving, stitching and three-

dimensional (3D) fiber architecture [16–20]. Among the various techniques to

improve interlaminar fracture toughness, stitching is more popular because it is

convenient and cost-effective [21]. Although it has been found that stitching

may cause degradation in some mechanical properties of the composite such as


4 Introduction

the tensile strength, it has been utilized in the manufacture of advanced 3D

composite materials with notable success since the early 1980s. So, some

research has been devoted to optimizing the stitching parameters [22–32].

Previous studies have investigated the effects of different parameters such as

stitching thread type, thread diameter, and stitch density on the performance of

the stitched synthetic fiber composites. It has been found that besides the stitch

density and type, the stitch distribution also plays an important role in

determining the interlaminar fracture toughness of the composites [33,34]. But,

unlike the synthetic fiber reinforced composites, the influence of stitching on

the properties of natural fiber composites is not as widely studied, and its

performance – when dealing with natural fiber composites – in view of stitch

density and material type is not that well understood [35].

1.3. Objectives of the research


The main purpose of this thesis is to investigate the effects of through-the-

thickness stitch reinforcements as well as preform architecture on the

mechanical properties, interlaminar fracture and impact behavior of the flax

fiber laminates impregnated by thermoset epoxy matrix. Epoxy/ flax fiber

composite systems with different stitch fiber areal fraction, stitch material and

in-plane fiber architecture (using available materials in the market for composite

application) were manufactured. The specific objectives of this research are to:

 Characterize mechanical properties including tensile and flexural

strength and stiffness, Mode I interlaminar fracture toughness and low-

velocity impact behavior of unstitched and stitched flax fiber laminates.


Chapter 1 5

 Understand the effects of through-the-thickness stitching on the

mechanical performance and fracture mechanism of flax fiber

composites when different stitch material and stitch areal densities are

used.

 Subject the flax fiber composites to low velocity impact to study the

effects of through-the-thickness stitching on the damage resistance,

impact energy absorption, and failure characteristics.

 Develop a numerical approach to model delamination crack propagation

of stitched flax fiber composites, and measure their interlaminar fracture

toughness through a virtual Double Cantilever Beam (DCB) test.

 Implement a numerical simulation to study the material behavior and

damage mechanisms at macroscopic scale of the woven flax fiber

composite under low-velocity impact loading.

The results of this study may contribute to a deep understanding of the

effectiveness of through-the-thickness stitching in enhancing the performance

of flax/epoxy composite laminates. It could also provide insights into the

material response of composites made of flax fibers through numerical

modeling in applications subjected to impact loads such as automotive segments.

1.4. Structure of the thesis


The thesis is organized as follows. A review of the state of the art on flax fibers

and their composites is presented in Chapter 2. Then, Chapter 3 describes the

material, experimental methodology, manufacturing technique and simulation

strategies used in this study. Chapter 4 presents the results and discussion of the
6 Introduction

tensile, flexural and DCB tests of various un/stitched flax fiber composites.

Based on the results and observations of Chapter 4, a numerical model is

developed in Chapter 5 to simulate delamination propagation in stitched flax

fiber composite. Chapter 6 is devoted to the experimental assessment of the low-

velocity impact behavior of the un/stitched composites. A numerical simulation

of the low-velocity impact test which was used to study and predict the material

behavior and damage propagation of the woven flax fiber composite is then

presented in Chapter 7. Finally, the conclusion of the study and our

recommendations for future work are presented in Chapter 8. The outline of this

thesis is schematically illustrated in Figure 1.1.


Chapter 1 7

Figure 1.1. Schematic outline of this thesis.


8 A Literature review

Chapter 2
A Literature review

2.1. Introduction
This chapter presents a brief review of natural fiber composites and their various

properties, with a particular focus on flax fiber composites. The constitution and

characterization of these composites are outlined and various methods to

improve their mechanical properties are described. The effects of fiber

architecture, fabrication and the interactions between the matrix and

reinforcement are then briefly discussed. This is followed by an overview of the

low-velocity impact behavior of laminated composites. Then, the effects of

through-the-thickness stitching on mechanical and impact properties of

composite laminates are reviewed. After reviewing latest studies on the

numerical modeling of natural fiber composites, this chapter ends with

highlighting the numerical methods reported for the stitched composite as well

as the failure analysis of woven fabric composites.


Chapter 2 9

2.2. Trend in using plant fibers to reinforce polymers


Recently, the use of plant fibers to reinforce the polymers have grown widely

in many applications thanks to their promising mechanical properties as well as

environmental and economic benefits compared to other types of available

natural fibers (Figure 2.1). Plant fibers own unique advantages of combining

good mechanical and vibration absorption (damping) properties with low

weight. Good acoustic and thermal insulation properties due to the hollow

structure of the plant cell (lumen) make them a very suitable material for interior

components, especially in automotive industries. Above all, they are renewable

(yearly based), sustainable and carbon dioxide (CO2) neutral, which make them

a desirable material while the environmental awareness is increasing [36–39].

It is worth mentioning that the term ‘natural fibers’ will be used to refer to plant

fibers onward in this study.

Classification of Fibers

Natural Synthetic

Plant Animal Mineral Synthetic


polymer
Natural
polymer
Metal Ceramic

Wood Fiber Non-Wood Sillk Wool Hair Asbestos (others)

Hard
Recycling
wood/Softw
woods Bast Leaf Seed/Fruit Stalk Grass, Grain
ood

Bamboo,
Pineapple, Cotton,
Flax, Hemp, Wheat, Bagasse,
Abaca Kapok,
Maize, Esparto,
Jute, Ramie, Henequen, Milkweed,
Brley, Rye, Sabei,
Kenaf Sisal, Coir, Oil
Oat, Rice Phragmites,
Banana palm
Communis

Figure 2.1. Classification of reinforcing fibers (reproduced from [40]).


10 A Literature review

Natural fibers have generally been considered as an alternative to synthetic

based glass fibers, as one of the widely used and market dominant

reinforcements in conventional composites [37,41]. Glass fibers have desirable

mechanical properties for applications in automotive, aerospace, marine and

construction. However, they are costly in terms of energy to produce and recycle.

It has been shown that plant fibers are the most appropriate and accessible

natural based alternative to glass fibers in many nonstructural applications when

the mechanical properties respect to density are taken into account.

Besides, natural fibers are environmentally superior to glass fibers. Their lower

density accompanying with good mechanical properties and environmental

advantageous make them a suitable reinforcement for polymers which can be

utilized in higher performance applications. Hence, extensive investigations

have been carried out to improve the properties of composites made of natural

fibers, with the purpose of extending the use of these fibers to structural

applications [4,1]. Table 2.1 shows the advantages of natural fibers over

synthetic glass fibers.

Table 2.1. General advantages of natural fibers over synthetic glass fiber [37].

Natural fibers Glass fibers


Density Low Twice of natural fibers
Cost Relatively Low Relatively high
Energy consumption Low High
Disposal Biodegradable Not biodegradable
Recyclability Yes No
Renewability Yes No
CO2 neutral Yes No
Health risk when inhaled No Yes
Abrasion to machines No Yes
Chapter 2 11

Consequently, the production of natural fibers for composites purposes is

growing rapidly. Global natural fiber composites market size reached $2.1B in

2010, with on average annual growth rate of 15% in last five years [42]. The

production of wood and natural fiber composites in the European Union

exceeded 352,000 tons in 2012, in which 92,000 tons belonged to natural fiber

composites. This was claimed to be an average of 10-15% of the total composite

market at that year [43]. It is also expected that their production will exceed

370,000 tons per year by 2020.

As of today, automotive interior parts are the most important application

segments for natural fiber composites, where the majority of them use short

fibers for injection molding and compression molding manufacturing. However,

by developing natural fibers for higher performance applications, their use will

be expanded and they will attract more attention in composite industries.

Figure 2.2 shows the performance playground and corresponding applications

for natural fiber (i.e. flax) composites, which depicts their potential to be used

from injection-molded components for mass production (moderate properties)


12 A Literature review

to structural composites by using specially designed fiber textiles like UD

prepregs.

350

300
UD flax
Composites
250 (45 vol%)

© CELC © Artengo © Museeuw Bikes


Tensile Strength (MPa)

200

150 Woven flax Composites


(45 vol%)

100 © Urge Bike © NoTox

50 Needle felt
Short fiber
(30 vol%) © AFT Plasturgie © Magis
0
0 5 10 15 20 25 30

Young's Modulus (GPa)

Figure 2.2. The performance and corresponding applications of composites made of natural
fiber (flax) in different fiber forms (reproduced form [44]).

However, there are still important issues that limit their future use, including

long-term performance and the ability to be able to predict performance during

service. Poor compatibility between fiber and matrix that and subsequently low

fracture toughness, sensitivity to temperature, moisture absorption, ultraviolet

that cause an initial crack and high variability in quality and homogeneity due

to the non-constant fiber properties fiber properties have been reported as the

main issues for composites made of natural fibers. Currently, many researches

are going on with the aim of finding new methods to fix these issues in order to

use natural fiber composites in load-bearing applications and replace the less

eco-friendly materials.
Chapter 2 13

2.3. Flax fiber composites


2.3.1. Flax fibers

Flax (Linum usitatissimum) is generally reported as one of the important natural

fibers in terms of good mechanical properties with a lower density which has

the potential to replace synthetic glass fibers in many applications [2,8,44–46].

This is mainly due to some promising characteristics in terms of microstructure,

chemical composition, and morphological properties of flax fiber in comparison

with other natural fibers.

Flax fiber can refer to either elementary or technical fiber. However, due to the

clear difference between the physical and mechanical properties, a distinction

should be explicitly made between them. Technical fibers are constituted from

10 to 40 elementary fibers, sticking together into a long thin fiber bundle

(Figure 2.3). The tensile strength and the initial modulus of elementary fibers

were found to be 1500-1800 MPa and 60-80 GPa, respectively. Whereas, these

properties for technical fibers were estimated to be about 800-1500 MPa and

55-75 GPa [47] due to load transferring between elementary fibers through

weak inter-fibrillar bonding, namely pectin, lignin, and hemicellulose.


14 A Literature review

Elementary fiber Technical fiber


(Also known as a fiber or cell) (Also known as a fiber bundle)

Figure 2.3. Schematic and SEM images of technical fibers versus elementary fibers (adapted
from [44]).

The chemical and physical characteristics of flax fibers have been the subject of

many studies [36,45,48–55]. Figure 2.4 shows the schematic of the hierarchical

structure of flax fiber from stem to microfibril. As can be seen, the flax fiber

cell wall, as for most plant fibers, is constituted of different layers, the primary

and secondary wall in which the secondary wall is subdivided into three layers

S1 to S3, forming together a multi-layered hollow (called lumen) tube

configuration. The S2 layer has the biggest dimension, almost 75-80% of the

cell wall, which is mainly constituted by highly crystalline cellulose microfibrils

fixed in helical structure with a characteristic spiral angle called microfibrillar

angle (MFA) along the fiber direction.


Chapter 2 15

Secondary wall S2 (75-80% of total thickness)


Microfibrillar angle (MFA) ≈10°

Figure 2.4. Hierarchal structure of flax fiber from stem to microfibril (figures were reproduced
from [44,54]).

In general, cellulose molecules among the other main chemical constituents, i.e.

hemicellulose, lignin, and pectin (for some plant fibers), are accounted as the

main structural constituent of a plant fiber cell which stick together to form the

microfibrils [54]. Such cellulose microfibrils and their arrangement (MFA),

which can be influenced by many factors such as climatic conditions, age, and
16 A Literature review

degradation process, determine the final mechanical properties of the fiber (i.e.

strength and stiffness). Table 2.2 gives an overview of the average mechanical

and physical properties reported in the literature for flax fibers as well as some

well-known natural reinforcing fibers. As can be seen, the cell wall of flax fiber

in comparison with other natural fibers contains a high level of cellulose content

(67-75% of weight) with a small MFA of 10%. These inimitable characteristics

of flax fibers yield higher mechanical properties in terms of strength and

stiffness [56].

It is also well known that the fiber length or aspect ratio has a great impact on

the mechanical propitiates of composites. Flax fibers have a relatively high

aspect ratio (fiber length/fiber diameter) which, depending on the fiber/matrix

interfacial bonding strength, is generally longer than the fiber critical length in

the composite. This advantage of flax fiber makes it possible to utilize the

maximum performance of fibers in terms of strength and stiffness due to a better

stress transfer between fiber and matrix.

Table 2.2. Chemical, morphological and mechanical properties of some natural fibers used for
reinforcing polymers [1,39,45,57–62].

Chemical and Physical Properties Mechanical Properties


Natural Aspect Tensile Tensile
fiber Density Cellulose MFA Elongation
ratio, strength Stiffness
(g/cm3) (wt%) (°) (%)
L/d (MPa) (GPa)

Flax 1.38-1.45 65-85 1700 10 800-1500 60-75 1.2-1.6


Hemp 1.4-1.6 70 1000 6 550-900 40-65 1.6
Cotton 1.5-1.6 90 900 - 300-600 12 3-8
Bamboo 1.4 26-43 1500 - 750-950 30-50 1.9
Jute 1.4-1.5 61-71 150 8 400-800 13-26 1.8
Ramie 1.44-1.5 68-76 7 - 500 44 2
Coir 1.1-1.3 32-43 25 30-49 220 6 15-25
Sisal 1.2-1.3 65 8 10-22 600-700 38 2-7
Chapter 2 17

The tensile performance of various natural fibers and E-glass are compared in

the Ashby plot of Figure 2.5. It can be clearly seen that glass fibers are always

superior in the absolute tensile strength and stiffness properties. Nevertheless,

when the factors of weight and cost are considered, the performance of flax

fibers is comparable with that of glass fibers. Shah [5] performed a comparative

study and showed that flax fibers (and bast fibers in general) offer high specific

properties which are comparable to glass due to the latter’s higher density.

Figure 2.5. Ashby plots comparing (a) absolute tensile properties, (b) tensile properties per
unit density for different natural fibers and E-glass [5].

2.3.2. Flax fiber reinforced composites

The performance of flax fiber reinforced composites have been studied by

numerous authors [2,7–9,38,47,50,52,63–69] and the effect of different factors

such as fiber processing and refinement, aging, fiber surface treatment, etc. have

been investigated. Apart from the effect of intrinsic properties of the flax fibers,

the performance of their composites can be influenced by other variables like

reinforcements’ forms (i.e. short fibers, long fibers, multiaxials and

unidirectionals), matrix type (i.e. thermoplastic and thermosetting) and


18 A Literature review

manufacturing techniques (i.e. injection moulding, compression moulding,

vacuum infusion, resin transfer moulding and prepregging). The effects of latter

factors, similar to those of conventional fiber composites, are significant in

determining the composite performance.

The absolute and specific tensile properties (with and without considering the

factor of weight) of different natural fiber composites in various composite type

in terms of matrix and reinforcement form are compared with that of glass fiber

composites in Figure 2.6. In this plot, flax fiber composites are always placed

at the most right side of the natural fiber composites’ bars as flax fibers usually

offer higher reinforcing potential among other natural fibers, Table 2.2.

Although the absolute properties of the natural fiber composites tend to be lower

than that of glass fiber composites, the specific properties lead to promising

values, indicating the possibility of replacing glass fiber by natural fiber

counterparts, particularly in the stiffness-critical applications.

However, bearing in mind the fibers’ vulnerability to thermal degradation,

moisture absorption and lacking information about the in-service behavior of

natural fiber composites, it is necessary to consider variables other than tensile

performance depending on the component function and objective [5,8].

Nevertheless, it is rather evident that for load-bearing applications, the random

mats must be replaced by aligned continuous flax fibers such as unidirectional

(UD) or textiles commonly used for conventional fibers [70]. This is mainly due

to the higher aspect ratio of flax fibers (Table 2.2) which can be regarded as

long fiber reinforcement if a proper fiber/matrix interfacial strength is achieved.

Therefore, researchers have looked into understanding and developing different


Chapter 2 19

fiber preform architectures in order to obtain higher mechanical performance of

composites [6,9,47,71].

Natural fiber composites Glass fiber composites

Figure 2.6. Comparison of the absolute and specific tensile properties of natural fiber
composites with E-glass composites [5].

2.3.3. Fiber Architecture

Four different types of flax fiber bundles are currently available for using in

composites: hackled flax, doubled flax, roving, and yarn [47]. Figure 2.7

illustrates the selection of flax fibers for different fiber preform architectures.

Similar to synthetic fibers, composites of unidirectional (UD) flax fibers have

the highest tensile properties in longitudinal direction [71] since the full

reinforcing potential of flax fibers is used. The UD fibers also allow having

different preform lay-ups with various stacking sequences and angle


20 A Literature review

combinations in order to provide desired properties in different orientations.

Nevertheless, they are not applicable for complex geometry parts as the fiber

bundles are detached easily during draping on mold, and this results in non-

uniform fiber dispersion in the composite. For that reason, other types of

reinforcement architectures, such as textiles, are generally used due to easier

composite processing.

Textile fiber reinforced composites can be used to produce many products with

a complex geometry and the desired performance. However, textile fiber

reinforced composites normally have lower mechanical properties due to having

lower fiber volume fraction, fiber misorientation as well as fiber damages

caused by twisting, weaving or braiding process.

Figure 2.7. Different flax fiber processing and selection of fiber preforms for composite [47].

In textile-based natural fiber composites, one of the challenging parts is to have

a textile produced by an appropriate yarn so as to use the maximum load-bearing


Chapter 2 21

capacity of the fibers to be suitable for engineering applications [6,72].

Although flax fiber yarns with a low level of twist lead to improved

impregnation of the polymer resin, they are very weak to be used in textile

processing machines, i.e. weaving, braiding or knitting. On the other hand,

increasing the twist level of flax fiber yarn results in misorientation of fibers,

and therefore reduction in mechanical properties of impregnated yarns

compared to UD composite [73,74]. Moreover, twisting can induce more

damage to technical and elementary fibers, and therefore degraded their

mechanical properties. Consequently, an optimum level of twist angle must be

used so that the maximum reinforcing potential of yarn (un-impregnated and

impregnated) can be achieved, Figure 2.8.

(a) (b)

Figure 2.8. (a) Effect of twist on yarn tensile strength and failure mechanism; (b) Tensile
strength of (●) long and (□) short flax fiber epoxy impregnated yarns as a function of twist
level [74].

Besides, further processing the flax fibers after scutching and combining step,

i.e. spinning and weaving to a textile, are the most expensive processes and
22 A Literature review

consume a large amount of energy [47,66]. In this regard, Shah et al. [74]

developed an analytical model to study the effects of the twist angle of flax yarn

on tensile properties based on the modified rule of mixture. Their developed

model was compared with experimental data, and good agreement was reported.

Goutianos et al. [6] developed high-performance natural fiber composites for

structural applications using continuous natural fiber reinforcements like UD

tapes or woven fabrics. Flax yarns with different twist angles were also tested

to find out the optimum twist in order to balance processability and mechanical

properties. Using the optimum twisted yarn, they produced various composite

laminates produced by fabrics with different weave structure and composite

processing technique and then characterized the mechanical properties. Their

study showed that continuous flax fiber reinforced composites were able to

compete with glass composites in terms of stiffness particularly from weight

reduction point of view.

It was observed that the tensile strength of impregnated yarn was improved by

decreasing the twist angle due to improved permeability, fewer damages to

technical fibers and better fiber alignment along the loading direction. In fact, a

twisted impregnated yarn is exactly similar to an off-axis UD composite in

which the tensile strength value is directly dependent to UD fibers orientation

angle. This features the necessity of using low twist yarn to utilize flax fiber’s

maximum performance.

Weagar [75] described a new technology of producing twistless yarns which are

developed by Composite Evolution Co. to optimize yarn stiffness. This


Chapter 2 23

company used a unique twistless technology to ensure a high degree of fiber

alignment, and improve the impregnation and performance.

The mechanical properties of natural fiber composites made out of long flax

fiber reinforcements have been the subject of several experimental studies. The

fiber preform architecture, stacking sequence and number of layers were found

to have a significant influence on the tensile and compressive properties of

composites [7]. For a wide range of fiber architectures such as woven, UD and

random orientation, Bensadoun et al. [8] have conducted static and fatigue

tension test to investigate the fatigue behavior of the flax fiber/epoxy

composites. They observed that the laminates with better fiber alignment such

as UD and Quasi-UD, not only show higher static strength and stiffness, but

also lead to the best fatigue characteristics (Figure 2.9).

(a) (b)

Figure 2.9. Comparison of static and post-fatigue tensile properties of flax–epoxy composite
after exposing to fatigue cycling up to 500,000 cycles, (a) the tensile modulus and (b) the
ultimate tensile strength [8].

Muralidhar [7] studied the effect of lay-up stacking sequence and number of

preform layers on the tensile and compressive properties of the plain woven flax

fabric/epoxy composite. Fiber volume fraction was found to be the main


24 A Literature review

parameter governing the tensile properties, whereas the increase in the number

of layers deteriorates their tensile strength.

The influences of yarn linear density, weave configuration and stacking

sequence of the woven flax fiber composites on the in-plane fracture toughness

have been studied by Liu et al. [9]. They performed a compact tension test to

evaluate the composite fracture toughness. It was found that the fracture

toughness of the composite is mainly dominated by the number of woven plies

as well as the linear density of the yarn (i.e. fiber volume fraction), rather than

the preform fiber architecture.

2.3.4. Thermoset resins

The role of the polymer matrix as the second constituent of composites is to

provide the composites’ shape, environmental tolerance, surface appearance,

overall durability and transferring the load to the reinforcing fibers as the

structural constituent of composites. There exist two major polymer types:

thermoset and thermoplastics. The difference between the chemical and

physical properties of these polymers determines the composite properties and

manufacturing process.

The thermoset polymers, unlike the thermoplastic ones, are liquid at room

temperature and molecularly cross-linked in a 3D network during curing which

then are not able to be molten. Due to these cross-link networks, the thermosets

generally yield better strength and stiffness compared to thermoplastics, and

therefore are a preferred matrix in fiber reinforced composites for structural

applications. Moreover, due to the low molecular mass of thermosets, their


Chapter 2 25

viscosity is low, and this provides better processability and impregnation of the

fibers by the resin using many common manufacturing techniques such as

vacuum infusion, resin transfer molding (RTM), pultrusion, autoclave, and hand

lay-up.

Thermoset resins also have a good chemical, thermal and creep resistance and

need a lower processing temperature. However, they are normally brittle and

have lower toughness [76]. Thanks to the low processing temperature and

desirable mechanical properties, thermoset resins are a suitable and preferable

matrix for flax fiber composites, particularly in high-performance applications

[5,36,54]. Utilizing long flax fibers in a thermoset matrix usually offers higher

mechanical properties, as shown in Figure 2.6.

2.3.5. Fiber-matrix bonding

One of the challenges of natural fiber reinforced composites is the poor

compatibility (adhesion) between fiber and matrix. The efficiency of stress

transfer from the matrix to the fibers, and consequently the composite’s stiffness

and strength can be reduced with a weak fiber/matrix interface by crack

initiation and propagation along the interface, i.e. debonding mechanism. Many

investigations have been conducted in order to improve the flax fiber/matrix

interface properties (i.e. interfacial bonding) and various physical and chemical

methods have been developed [2,69,77,78]. These methods are summarized in

Figure 2.10.
26 A Literature review

Fiber Treatment

Physical method Chemical method

Plasma Corona Thermo Enzyme Maleated Alkaline Silane


Acetylation
treatment treatment treatment treatment coupling treatment treatment

Figure 2.10. Classification of natural fiber treatment methods [1].

The physical treatments change the surface properties of fibers and thus improve

the mechanical bonding between the fiber and the matrix, whereas the chemical

treatments use chemical methods and coupling agents for enhancing the

fiber/matrix adhesion. For example, Huo et al. [79] investigated the effects of

various chemical modifications on the mechanical performance of

flax/vinylester and compared with E-glass/ vinylester. The study showed that

the specific properties of chemically treated UD flax fiber composites are higher

those of E-glass composites. Also, a 30% increase in the interfacial and the

interlaminar shear strength of Acrylic acid treated flax fiber composite was

observed.

2.3.6. Composite processing

Polymer matrix composites with synthetic fiber reinforcements are processed

and manufactured by numerous methods such as hand lay-up, prepreg, liquid

compression molding, pultrusion, filament winding, autoclave processing and

liquid composite molding (e.g. resin transfer molding, vacuum infusion, and

resin infusion) techniques [76]. The majority of these manufacturing methods

can be utilized in order to process natural fiber composites. Nevertheless,


Chapter 2 27

because of specific characteristics of the plant fibers, some parameters should

be controlled to have a successful manufacturing process. Moisture, thermal

transition temperatures, exotherm2, volatile components, and rheology3 are the

key parameters of processing natural fiber composites [36]. Furthermore, due

to surface roughness, morphology and lower degree of fiber alignment, natural

fiber preforms normally show higher resistance against compressing, and

thereby a higher clamping force (of the mold) is needed. This means that low

fiber volume fractions are achieved if the manufacturing techniques like hand

lay-up and vacuum infusion are used, in which the ultimate compaction pressure

is only 1 Atm [80]. Figure 11 presents Ashby plots which compare the tensile

strength and stiffness of various natural fiber composites with different fiber

architectures as well as different manufacturing techniques.

Figure 2.11. The performance of various natural fiber composites manufactured using
different techniques [46].

2 The heat generated during polymerization or cross-linking (curing) of polymer resin.


3 The science of flow and deformation of matter.
28 A Literature review

2.4. Impact behavior of laminated composites


2.4.1. Composite laminates under low-velocity impact

The transverse performance of composite laminates is a concern when they are

known to suffer low-velocity incidents such as tool drops, stone or debris strike,

etc. Typically, incidents occurring below 10 m/s are considered as low-velocity

impact [81]. The damage size and type caused by an impact generally determine

the residual properties of the structure made of composite laminates [82–85].

Hence, the low-velocity transverse impact behavior of composite laminates has

always been one of the main design considerations in many applications [13,14].

Low-velocity impact behavior (such as dropping weight) analysis for composite

materials generally consists of three major parts: i. Damage resistance, ii.

Damage tolerance and iii. Energy absorption [10,85–88].

i. Damage resistance is defined as the ability of a material to resist

permanent changes due to a loading event beyond the design load. The

damage resistance of composite laminates is normally assessed after an

impact event by measuring the damaged area using Non-Destructive

Inspection (NDI) and visual inspection.

ii. Damage tolerance refers to the ability of a material to carry the load after

a permanent change taking place. Once a laminated composite has been

impacted, the residual compression strength is usually of most interest

due to the presence of nonvisible delamination as the main failure mode

under low-velocity impact. Delamination in laminates can reduce the

bending stiffness by dividing a laminate into sub-laminates. A

delaminated laminate is also susceptible to buckling under compressive


Chapter 2 29

loading. Compression after impact (CAI) and Flexure after impact (FAI)

tests are commonly used to evaluate and characterize the residual

properties and functionality of composite laminates.

iii. Energy absorption is characterized by the amount of impact energy

which is dissipated by the material through the development of internal

damages. The absorbed energy in low-velocity impact can be

determined by calculating the area under the force-displacement curve.

The maximum capacity of a laminate to absorb impact energy can be

determined when the laminate is penetrated – i.e. full perforation impact

test.

Depending on the application requirement, the aforementioned impact

properties, as design parameters, should be determined. For instance, for a

product with energy absorption purposes, the ultimate perforation energy should

be maximized [86,89]. Literature suggests that the fiber strength and the fiber

volume fraction have a strong effect on the perforation characteristics of

composites, whereas matrix type, fiber orientation and architecture, and

thickness only play a secondary role [47,85]. The effects of these factors can be

even more complex when a lower-strength reinforcing fiber (e.g. from natural

origin) is used.

The load and energy histories of low-velocity impact can yield important

information about the damage initiation and propagation mechanisms. Typical

force and energy histories for drop-weight impact tests of non-penetrated and

penetrated laminates are given in Figure 2.12. The characteristic values of


30 A Literature review

damage initiation force (P1), maximum force (Pmax), maximum absorbed energy

and contact duration are achievable from these graphs.

(a)

Pmax

Energy recovered due


to elastic response
EAbsorbed
P1 (absorbed energy )

(Damage initiation)

Energy dissipated by damage


development
(b)

Pmax

P1
EAbsorbed
(Perforation energy )

Friction force of impactor

Figure 2.12. Typical impact load–time and impact energy–time histories of (a) non-penetrated
and (b) penetrated laminate [87].
Chapter 2 31

The impact damage mechanism of composite laminates is a very complex

process, a combination of delamination, matrix crack, fiber shear-out, fiber

breakage, and surface buckling, etc. [87]. The matrix cracks are believed to be

the primary impact damage mode in all types of composites [90]. However, it

was shown that formation of matrix micro-crack does not lead to a noticeable

reduction in stiffness of the laminates. However, the local and global stiffness

of laminates are strongly affected by initiation of delaminations and fiber

breakages. Hence, delamination threshold loads, which generally occurs prior

to fiber breakage, is identifiable from the first sudden load drop in the impact

load-time history (i.e. P1) [90,91]. The first delamination normally appears

between the top plies for stiff laminates, and forms between bottom plies when

the laminate is thin or flexible. A typical impact damage mechanism for

laminated composite is depicted in Figure 2.13.

Figure 2.13. Schematic illustration of a typical impact damage mode for composite
laminate [87].

Many factors influence how much damage is incurred by a laminate from a

foreign object impact event. Impact velocity and energy, laminate thickness,

boundary conditions, the shape of the striker's head, etc. are of the most
32 A Literature review

important external factors for a given laminate. ISO 6603 and ASTM D7136

[81,92] defined standard procedures for measuring the impact behavior of a

synthetic fiber reinforced composite through a drop-weight impact event.

2.4.2. Low-velocity impact behavior of natural fiber composites

With the purpose of extending the use of natural fibers to higher performance

applications, the low-velocity impact response and energy absorption behavior

of their composites, using drop-weight impact test, have been the subject of

several investigations [10–12,15,47,93–100].

The effect of fiber preform architecture and laminate thickness on the low-

velocity impact performance of jute/methacrylated soybean oil were studied by

Dhakal et al. [99]. Both the fiber orientation and laminate thickness exhibited a

significant influence on the impact resistance and peak load. It was also

observed that improving the fiber/matrix interface through treatment of jute

fibers led to improved impact resistance of the composite.

Impact behavior of UD, cross-ply [0/90°] and mat hemp preforms with epoxy

resin was investigated by means of drop weight test by Santulli et al. [94]. It

was observed, although the mechanical properties (i.e. tensile, modulus and

flexural strength) owned higher for UD and cross-ply laminates, the mat fiber

laminate showed better impact resistance in terms of penetration energy and

damage area. It was concluded that more complex and tailored fiber architecture

of composite would likely increase the impact performance of the composites.

Vasconcellos et al. [97] studied the post-impact fatigue behavior of woven

fabric hemp/epoxy composite through cyclic tensile tests and the acoustic
Chapter 2 33

emission monitoring to analyze damage mechanisms. Their results showed that

while the residual tensile strength decreases with increase in the impact energy,

the elastic modulus remains almost unchanged. Fiber/matrix interface and

matrix micro-cracking were observed as the main fracture mechanism caused

by non-perforation impact test.

Caprino et al. [100] studied the effects of various preform architecture including

UD, woven fabric and cross-ply (0/90)s, as well as the laminate thickness on the

behavior of hemp/epoxy composites subjected to a low-velocity impact. They

also validated the existing semi-empirical models and analysis methods – i.e.

Hertzian contact law and exponential indentation law – valid for synthetic fiber

laminates and found that these models were not able to interpret the impact

behavior of bio-laminates. Moreover, unlike the conventional composites, a

very limited delaminated area was found at the impact point for all the laminates.

In another study [11], the impact and post-impact behavior of flax/epoxy with

two different lay-up sequences ([0/90/45/-45]2s and [90/0/-45/45]2s) were

investigated by means of drop-weight impact and compression after impact tests,

respectively. Similar to the previous study, a linear correlation was found

between impact energy, absorbed energy and maximum bending deflection.

Also, a maximum loss of 30% in compressive strength was observed for the

2.85 mm thickness laminates impacted by 10J. Comparing the force-deflection,

absorbed energy and damage size of impact and indentation tests showed that

the flax/epoxy composites response under transverse loading condition were not

rate-dependent when the loading rate was below 2.98m/s.


34 A Literature review

Petrucci et al. [12] studied the impact and post impact performance of

natural/glass fiber hybrid composites through drop weight impact and FAI tests,

respectively. They manufactured various laminates through stacking Hemp,

basalt, flax and glass biaxial textile and processed using vacuum infusion.

Enhanced flexural modulus was observed for all the hybrid composites in the

presence of glass fibers, and among them, flax/basalt/glass hybrid composite

had the highest flexural performance before and after impact. This was

attributed to the better interaction between glass and flax fibers. In terms of

impact properties, although the natural/glass fiber hybrid composites had higher

impact resistance, pure natural fiber hybrid composites (flax/hemp) offered

better penetration energy.

With the purpose of characterizing ballistic properties of natural fiber reinforced

composites, a ballistic impact test conducted on plain woven flax, Hemp and

Jute fabric reinforced composites as well as composite/mild steel hybrid [101].

It was observed that the flax composite absorbed the impact energy slightly

more than mild steel and other composites. However, composite/steel hybrid

(steel backing) showed clearly better performance. A similar study was also

carried out using an aluminum plate for backing by Kuan et al. [102] and a

significant enhancement in the tensile and impact properties of the laminates

was reported.
Chapter 2 35

2.5. Through-the-thickness stitched composites


Composite materials mainly suffer from weak interlaminar properties due to

lack of out-of-plane reinforcement. Hence, layer-by-layer separation, which is

known as delamination, is usually the weakest failure mode in laminated

composites. So far, various techniques have been introduced to enhance the

interlaminar strength of composites; for example engineering the interface

adhesion between fiber and matrix, increasing the toughness of matrix,

interleaving and three-dimensional (3D) fiber architecture [16–20]. Three-

dimensional textile technologies (3D fiber architecture) such as 3D weaving,

braiding, knitting, and stitching have been developed and widely utilized in

order to provide through-the-thickness reinforcement into the composites and

improve the interlaminar properties [19]. Among the different 3D fiber

reinforcements, the stitching process became more popular by reason of

convenience and cost-effectiveness of this method [21].

Since the early 1980s, stitching has been utilized in the manufacture of advanced

3D composite material with notable success, Figure 2.14. The aircraft industry

was the first one to start an investigation on the stitching of composites to

improve properties of fiber reinforced plastic (FRP) joints. They investigated

the feasibility of stitching wing-to-spar and single-lab joints of composite

components to increase the failure strength and reduce the probability of sudden

catastrophic failure. The result of these investigations acknowledged the

benefits of stitched joints in comparison to conventional joining techniques such

as adhesive bonding and co-curing [21].


36 A Literature review

(a) (b)

Figure 2.14. Example of (a) complex preform manufacturing through stitching and (b)
commercially available robotic stitching machine (Altin Nahtechnik GmbH) [21].

Two main characteristics of the stitching process can be considered as the main

reason for its attractiveness and popularity. Firstly, stitching is a cost-effective

method for joining stacked fabric plies along their edges to make the preform

easier to handle before resin infusion or liquid molding. The second advantage

of stitching is that it can improve the delamination resistance and impact

damage tolerance of composites [23,82,103–105]. Consequently, a considerable

amount of research has been devoted to evaluating the different properties of

stitched composite laminates.

Despite these advantages, damage to the preform induced by the stitching

process is the main drawback of this technique, where it can deteriorate the

properties of the composite and therefore limit its use. Stitching can cause

different types of damage to fiber preforms. These damages are: fiber breakage,

fiber misalignment, fiber crimping, resin-rich areas, stitch distortions,

compaction, and matrix micro-cracking, in which fiber breakage, misalignment


Chapter 2 37

and crimping (Figure 2.15) are the most common types of damages in the

stitched composites [21,105–107].

(a) (b) (c)

Figure 2.15. Examples of stitching damages; (a) in-plane fiber breakage, (b) local
misalignment of fibers around a stitch and (c) schematic of fiber crimping [21,108].

In addition to preform damage, the stitch fibers also can be damaged by twisting,

bending, sliding and looping during stitching process, particularly when brittle

fibers like carbon are used as stitching thread [109,110]. In this regard,

Weinberg et al. [111,112] invented a sewing machine needle and a sewing

mechanism which is useful for stitching and joining fabrics with a brittle thread

(Figure 2.16).

Figure 2.16. Diagrammatic side view of the sewing machine and its particular needle
developed for the brittle thread (reproduced from [112]).
38 A Literature review

There is variety of stitching patterns which are used in the composite industry,

and of those the most popular patterns are lock stitch, modified lock stitch, and

chain stitch, as shown in Figure 2.17. The lock stitch pattern is mainly utilized

in the apparel industry for aesthetic purposes as the loop of needle and bobbin

threads is hidden within the preform. However, the looping process between the

needle and bobbin threads creates disruption to the preform. To avoid these

defects inside the laminate, the modified lock stitch is normally used for

structural purposes in which the loops of the bobbin and needle yarn are placed

at the outer surface of the preform [21,27].

(a) (b) (c)

Figure 2.17. Schematic illustration of (a) lock stitch, (b) modified lock stitch and (c) chain
stitch.

2.5.1. Effect of stitching on mechanical properties and toughness

of composites

2.5.1.1. Interlaminar fracture toughness

The growing use of stitched composites in various load-bearing applications

required an in-depth understanding of their mechanical properties and failure

mechanisms. Consequently, a considerable amount of researches have been

devoted to evaluating and understanding the effect of stitching on different

properties of composites. Interlaminar crack growth (i.e. delamination) of

stitched graphite/epoxy laminate with Kevlar thread, was initially investigated


Chapter 2 39

by Tan et al. [113] in the early 1980s. It was experimentally shown that through

the thickness stitching effectively arrested delamination progression. In addition,

it was revealed that stitching had different effects on laminates’ ultimate

strength.

Numerous studies have experimentally investigated the effects of different

parameters such as stitching thread type and stitch density on the performance

of the stitched synthetic fiber composites [24,26–28,33,34,108,114–116].

The influence of stitch distribution and in-plane fiber orientation on interlaminar

fracture toughness of stitched glass/polyester composites was investigated by

Solaimurugan et al. [108]. It was shown that it improves with an increase in

fiber orientation angle. Figure 2.18 compares the load versus crack opening

displacement (COD) and delamination mechanism – stable against stick-slip

propagation – of the unstitched and stitched laminates. The fractured and pulled-

out stitch fibers can be easily seen from Figure 2.18b.


40 A Literature review

(a) (b)

(c)

Figure 2.18. DCB test of (a) unstitched and (b) stitched specimen; (c) Load versus COD
curves of unstitched and stitched composites with similar fiber orientation [108].

Tan et al. [27] suggested that the stitch fracture process in the stitched carbon

fiber composites follows the steps shown schematically in Figure 2.19.

(a) (b) (c) (d) (e)

Figure 2.19. Stitch fracture process: (a) Initial stitch; (b) Interfacial debonding; (c) Slack
absorption; (d) Stitch fracture and (e) Frictional pull-out [27].

Literature has also shown that, apart from stitch density which is a key factor in

delamination suppression, the Mode I interlaminar fracture toughness of

stitched laminates is influenced by other factors like stitch fiber type and
Chapter 2 41

thickness, stitch distribution, in-plane fiber architecture, fiber orientation

[27,103,108,115,117]. For example, Wood et al. [117] conducted an

experimental study on interlaminar fracture toughness of Vectran stitched

CFRP laminates and revealed that there exist significant variations in the

interlaminar fracture toughness of stitched specimens with identical stitch

densities but different stitch distributions. Similar results have been also

reported by many authors [22,26,28,33,115,118].

2.5.1.2. In-plane mechanical properties

The effect of stitching on the in-plane properties of the composites is

unavoidable. Several investigations studied the tensile, compressive, flexure,

interlaminar shear and fatigue properties of stitched composites. In some cases,

they contradict each other’s results [23,24,26,32,105,106,114,116,119–124].

The majority of these studies indicate a reduction in the tension, compression

and flexural properties of the stitched laminates in comparison with their

equivalent unstitched counterparts [106,120,123]. However, there are some

studies that exhibit no change or a modest improvement to their mechanical

properties [32,105].

For example, Kang et al. [105] compared the tensile modulus and strength of

two stitched composites under identical conditions. Their results showed an

apparent contradiction, as illustrated in Figure 2.20. The increase in stitch

density results a continuous reduction in the tensile strength and modulus of the

S2 glass/polyester, whereas the Kevlar/PVB-Phenol shows an erratic increase

in the modulus with the same stitch density, while the strength remains almost
42 A Literature review

unchanged. Similar contradictions were also observed for the flexure and

compression properties of the stitched composites.

(a) (b)

Figure 2.20. Effect of increasing stitch density on the tensile strength and modulus of chain
stitched (a) S2 glass/polyester and (b) Kevlar/PVB-Phenol composites [105].

Mouritz et al. [32] conducted a comparison study on the substantial published

data for stitched in order to quantify the advantages and disadvantages of

through-the-thickness reinforcement for in-plane mechanical properties of

composites. Figure 2.21a-d show the effect of stitch density on the different

mechanical properties of various conventional stitched composites. Comparing

the results showed that stitching could either lead to improvement or reduction

in the tension, compression, flexural and interlaminar shear properties, usually

by less than 20%.

The reduction in the tensile strength of stitched composites is attributable to

geometrical stitch-induced defects such as waviness and crimping of in-plane

fibers, and fiber breakage during the stitching process [25,32]. These

mechanisms were also suggested for the decrease in the other mechanical

properties (i.e. compression and flexural) of the stitched composites. While, no


Chapter 2 43

certain mechanism has ever been directly observed for the increase in the tensile

strength. In the absence of a decisive inference, a plausible general mechanism

for the increase in the in-plane properties of stitched composites is an increase

in the in-plane fiber volume fraction due to compaction of the stitched preforms

by the tension in the stitch fibers during processing.

Mouritz et al. [25] suggest that the increase in the in-plane fiber volume fraction

is almost the main reason for the increase in the elastic modulus of stitched

composites compared to the equivalent 2D laminate. Moreover, the ability of

stitching to suppress delamination propagation is another plausible mechanism

for the improvement in the properties like the flexural strength of the stitched

composites.

From reviewing the published data on the stitched composites, it can be said

that predicting the effects of stitching on the tensile properties is not

straightforward and it is influenced by many parameters other than stitching

density. Fiber preform architecture, the thickness of the laminate, stitch

distribution, stitch thread diameter, stitch type and matrix type are some of the

most important influencing parameters on the performance of stitching in terms

of tensile properties of stitched composites [106,109,125–127].


44 A Literature review

(a) (b)

(c) (d)

Figure 2.21. Effect of z-binder content (stitch density) on the mechanical properties of various
synthetic fiber composites; (a) normalized tensile strength, (b) normalized tensile modulus, (c)
normalized flexural strength, (d) normalized flexural modulus (The normalized properties,
defined as the ratio of properties of stitched composites to that of the equivalent unstitched
laminate) [32].

Unlike the synthetic fiber reinforced composites, the influence of stitching on

the properties of natural fiber composites is not as frequently reported. Rong et

al. [35] carried out an investigation on the factors that influence in-plane

mechanical responses and Mode I interlaminar fracture toughness of

unidirectional sisal/epoxy laminates stitched by Nylon and Kevlar threads. It

was found that tensile and flexural properties were not significantly affected by

stitching, while the delamination resistance was improved via expanding the
Chapter 2 45

fiber bridging zone. To our knowledge, this is the only work which has done

experimentally on the effect of stitching on the natural fiber composite.

2.5.2. Effect of stitching on impact behavior of composites

In general, an important failure mode in composite laminates subjected to

transverse loading is the interlaminar failure or delamination. Stitching as

known to be a very convenient and cost-effect method in the suppression of

delamination propagation. In this regards, the impact performance of stitched

synthetic fiber composites has been investigated by many authors

[26,31,82,106,128–134].

Mouritz et al. [135] conducted an experimental study to compare the damage

resistance of the unstitched and stitched glass fiber laminates using Kevlar yarn

under various impact energies. Delamination area, residual flexural and shear

strength were measured under condition of increasing energy and number of

repeated impact tests to evaluate the damage. They observed that although the

Mode I interlaminar GIC was improved by stitching, the interlaminar shear

strength of stitched laminates decreased owing to stitch-induced damages in the

homogeneous resin-rich region between fiber plies. A slightly larger

delaminated area, therefore, was observed for stitched laminates under both

single and repeated impacts. In contrast to these results, a significant reduction

(up to 40%) in the delamination area was reported for impact on Kevlar-stitched

carbon/epoxy composites by [134].

Their results showed that the compressive strength and modulus decreased by

stitching, nevertheless, the compression properties after impact slightly


46 A Literature review

enhanced for the stitched laminates. These improvements are only observed for

the impact energy lower than 35J, and above this value, the effects of stitching

were negligible.

Tan et al. [14,26,129,132] conducted a comprehensive study on the effect of

stitching on the performance of the carbon fiber composites subjected to impact

loading. Based on an extensive series of experiments using specimens of

different laminate thicknesses, stitch densities and a range of impact energy

levels, it was concluded that out of three main stages of damage (i.e. damage

initiation, damage propagation and final damage failure), stitches only reduce

the damage initiation threshold. This is due to the presence of weak resin-rich

regions and stress concentration around stitch threads which act as crack

initiation sites. They observed that matrix cracks induced by stitches were

joined between stitch loops, resulting in an easier formation of the crack,

particularly in densely stitched composites. However, during damage

propagation, stitching plays a significant role in suppression of impact-induced

delamination by effectively bridging interlaminar cracks and arresting crack

growth. Therefore, stitched laminates had very smaller delamination region in

comparison with the unstitched counterparts, resulting in higher CAI failure

load and displacement (damage tolerance) of the stitched laminates. Moreover,

it was found that the final failure load improved with increasing density of

stitching, while the failure mechanism changed from delamination for

unstitched laminates to in-plane fiber breakage and matrix crushing for densely

stitched laminates.
Chapter 2 47

The use of flax fibers through hybridization with synthetic composites to

improve energy absorption properties has been the subject of some studies

[36,136–138].

Ghafari-Namini et al. [136] investigated the effects of flax yarn stitching on the

energy absorption and crashworthy behavior of hybrid composite

(Glass/Carbon) box structures. The result of DCB test of stitched and unstitched

composite having a similar lay-up showed that flax yarn stitching improved the

interlaminar fracture toughness of the composite. It was observed that the

lamina bending crushing fracture mechanism of unstitched composite box

changed to brittle fracture mode during crushing process. They concluded that

the improvement of the crush force efficiency and energy absorption

performance of the crush box depends on the location of stitched area in the box.

In another study, the performance of flax yarn stitched single lap joint of

(carbon/glass) hybrid composite beam under impact loading was investigated

[36]. It was found that composite beam joints with flax yarn stitches exhibited

higher impact resistance compared to adhesive bonded joints through modifying

and retarding the failure mode.

2.6. Numerical analysis of stitched and woven composites


Numerical simulation can provide a better understanding of the behavior of

material where it is difficult or impossible to obtain it experimentally. The

numerical simulation can also be used to conduct virtual testing for different

characteristics of the material in a shorter time and with less expenses.


48 A Literature review

In recent decades, the numerical tools to mirror the complex failure mechanisms

of the advanced fibrous composites has achieved tremendous success due to

high demand for their use as primary structural material in many applications.

One of the most powerful numerical methods for the structural analysis of

advanced composites is the finite element (FE) method, which enables to

capture the damage initiation and propagation mechanisms of synthetic fiber

composites using different failure theories. Some of these theories, despite

many differences in the micro-structure and properties, may be used to predict

the behavior of composite structures made of aligned discontinuous natural

fibers.

The following sections present a literature review focusing on the prediction of

interlaminar fracture toughness of stitched composites through DCB test

simulation as well as failure analysis of woven composites.

2.6.1. Numerical modeling for natural fiber composites

Numerical models to predict the mechanical properties of natural fiber

composites not only provide a better understanding of their behavior, but also

can help to extend their use to engineering applications. Recently, a few

research works have been dedicated to developing numerical models to predict

the mechanical behavior of natural fibers and their composites. These models

mainly focused on microscale modeling of natural fiber composites.

Modniks et al. [139] used the orientation averaging method to predict the tensile

stress–strain behavior of short fiber flax/polypropylene composites by means of

FEM simulation of a unit cell (Figure 2.22a). In another study, Sliseris et al.
Chapter 2 49

[140] developed a microscale model to simulate the mechanical properties of

short fiber-reinforced and woven fabric flax/epoxy composites (Figure 2.22b).

Their micromechanical model was based on the random generation of short

fibers and their bundles as well as fiber defects within a unit cell to study by

FEM the effect of the orientation of short fibers and their defects (e.g. kink band)

on the tensile properties of the composite.

Mattrand et al. [141] used a numerical approach to generate randomly the cross-

section geometry of flax fibers inside the composite, which can contribute to the

micromechanical modeling of natural fiber composites.

Beakou et al. [142] numerically studied the tensile properties of a flax fiber

bundle, taking into account the fiber discontinuity and the cohesive interface

between elementary fibers within the fiber bundle.

(a) (b)

Figure 2.22. (a) Schematic of the unit cell used in [139]; (b) Microstructures of short flax
fibers in matrix with different orientation coefficient [140].

These models, which are based on the micro-scale level modeling of the fiber

configuration within composites, are helpful in understanding the effect of


50 A Literature review

various intrinsic characteristics of natural fibers on the mechanical and fracture

properties of their composites. However, the implementation of these models

on macro-scale is computationally very expensive.

In addition to the micromechanical models which are based on the

reproducibility of a unit cell with FEM, a few constitutive models have

developed to consider the viscoelastic and viscoplastic behavior of natural fiber

composites [143,144]. Rubio-López et al. [144] proposed a constitutive model

to predict the viscoplastic behavior of PLA based composites reinforced with

three different woven fibers (flax, jute, and cotton). They showed that the

proposed model was able to predict successfully the behavior of bio-composites

loaded at different strain rates.

2.6.2. Predicting interlaminar fracture toughness of stitched

composites

It has been well known that stitching is an effective way to suppress

delamination growth in laminated composites. Besides the experimental

investigations presented in Section 2.5, a considerable amount of numerical

studies has been devoted to modeling the interlaminar failure of the stitched

synthetic composite laminates [22,23,27,33,34,110,117,145–147].

Mai et al. [110] proposed two micro-mechanics based models to study the

effects of stitching on delamination growth of a laminate in double cantilever

beam (DCB) specimen. In their first model, it was assumed that the stitches are

not interconnected at laminate surfaces since in most cases the top and the
Chapter 2 51

bottom surfaces of the stitched laminates are ground off to remove surface

waviness created by the stitch loops. Hence, each stitch was considered as an

independent through-thickness reinforcement, as shown in Figure 2.23a and b.

In the second model, the effect of interconnected stitches which had been

ignored in the first model was taken into account (Figure 2.23c). They assumed

that the matrix-stitch thread bond is completely frictional and constant. In

addition, the elastic bond strength and the effect of fiber abrasion or matrix

crumbling during stitch thread stretching and pull-out were neglected. It was

also assumed that the breakage of the stitch is brittle and its properties are

uniform over the length, and therefore stitches fracture at the delamination face

where the bridging force is maximum. The results of their analysis provided a

good understanding of the influence of stitch thread size on the delamination

resistance.

(a) (b) (c)

Figure 2.23. Illustration of an embedded stitch fiber (a) during elastic stretching and (b) after
slippage of embedded end without interconnected stitched effect; (c) interconnected stitched
fiber during elastic stretching [110].

An analytical model using a continuous cohesive spring with linear elastic

behavior and assuming brittle fracture was presented by Sankar et al. [148] to
52 A Literature review

simulate crack propagation in stitched composites. The authors concluded that

linear modeling is not adequate to predict the interlaminar fracture toughness of

the laminates stitched by a strong thread (i.e. Kevlar), where the inelastic

behavior of the stitching thread plays a significant role in increasing the

toughness.

Chen et al. [149] investigated the use of 2D solid elements and 3D shell

elements to model the parent laminate and bar elements to model the stitches to

determine the effective Mode I and Mode II fracture toughness of the stitched

composites based on the J-integral method. Their study showed that 3D model

is necessary to determine the accurate stress field as well as the stitch

interactions with the parent laminate.

Sun et al. [33] used a 2-node nonlinear rod element with the micromechanical

stitching models introduced in [110], and studied the effect of stitch distribution

on improving the delamination resistance by means of the virtual crack closure

technique (VCCT). In this model, the stitch elements were located between 2

sub-laminates in which their initial length was zero (before any load was carried

by the stitches) and it was assumed that only tension loads could be carried by

the stitch elements.

A comprehensive experimental and numerical study on the effect of stitch

distribution has been conducted by Wood et al. [117]. They employed the 2-

node beam element used by Sun [33] and the VCCT to simulate the DCB test

of the stitched composite. Their studies revealed that, regardless of stitch density,

the stitch distribution has a significant effect on the Mode I critical strain energy

release rate (G IC).


Chapter 2 53

Iwahori et al. [23] and Tan et al. [27] developed a 2-D finite element model to

simulate the delamination propagation of stitched CFRP for the DCB test by

using a 3-node rod element to represent a stitch thread. The authors also

developed a novel test method, named as interlaminar tension test, to understand

the progressive damage behavior of a single stitch thread – i.e. interfacial

debonding, slack absorption, stitch fracture, and frictional pull-out – as it is

loaded in tension (Figure 2.24). The load–displacement curve from the

interlaminar tension test was simplified and used in defining the material

modeling of the stitch element in the FEM simulation (Figure 2.25). The model

was able to provide a good prediction of the experimental load-displacement

curves and Mode I critical strain energy release rates, GIC.

(a) (b) (c)

Figure 2.24. (a) Interlaminar tension test specimen proposed by [27]; (b) fractured stitch after
interlaminar tension test; (c) fractured stitches at the surface of the test specimen after DCB
test [23].
54 A Literature review

Figure 2.25. Modeling of z-fiber mechanical properties [23].

Generally, it is observed that the interlaminar failure of long fiber reinforced

composites happens in the presence of in-plane fiber bridging which leads to an

R-curve development in the fracture energy response [150–154]. The bridging

traction of in-plane fibers across the delamination plane of the stitched laminates

using high strength stitch threads is negligible compared to the bridging traction

of the out-of-plane stitch fibers. However, the latter has to be taken into

consideration when preform stitching is conducted using lower strength

materials, like natural fibers. For modeling delamination in the presence of fiber

bridging, several authors have shown that the traction-separation law (cohesive

law) with non-linear softening can be used to model the R-curve response as

well as the enlarged process zone associated with crack bridging [150–153,155].

Airoldi et al. [151] proposed a semi-analytical approach to extract the

parameters of the cohesive law with bilinear softening (trilinear cohesive law)

– which is being obtained from the superposition of two bilinear cohesive laws

– from the experimental R-curves in order to model the delamination crack

propagation in the presence of fiber bridging. It was shown through a virtual


Chapter 2 55

DCB test, that the model was able to reproduce well the force-displacement as

well as the R-curve response.

2.6.3. Failure analysis of woven composites under impact loads

As of today, a variety of models have been proposed to successfully predict

various effective properties of woven composites using either simplified

macroscale models [156–161] or accurate models based on multiscale

approaches [162–166]. The heterogeneity of woven composites is generally

defined at three length scales based on their geometrical characteristics as

microscale (fiber), mesoscale (tow), and macroscale (laminate) [166]. The

multiscale modeling of woven composites is able to capture the effects of the

interlacing and undulation of the fiber tows at the mesoscale, the diffused

fracturing processes at the microscale within the fiber tows as well as depict the

effective behavior of composite plies at macroscale. This is possible by virtue

of homogenization theory which is applied between the different length scales.

In the multiscale models, the heterogeneous material is substituted by an

equivalent homogenous medium in the macroscale model using a proper

homogenization approach. To capture the micro and meso-mechanisms of

damage, stress states are computed at micro and mesoscale models. The stress

amplification factors, which are determined from the stress concentrations

analysis at meso and microscale FE models with the detailed woven architecture

of fiber tows, are used to transfer the stress states between different length scales.

This approach has been studied by numerous authors with the aim of integrating

different modeling tools within a single framework to depict the mechanical

behavior of woven composites. For instance, Lomov et al. [166] proposed a


56 A Literature review

comprehensive algorithm of mesomechanical failure analysis including a

preprocessing method for modeling of woven geometry. A multiscale approach

proposed by Ernst et al. [167] in which micro and mesoscale FE models were

used to derive the nonlinear behavior and the mechanical properties of the

woven laminate (the larger entity). Mao et al. [165] proposed a multiscale

modeling approach starting from the micromechanics to predict mechanical

behavior until fracture of woven composites.

Although, due to this heterogeneous structure, the micro and mesoscale

modeling are necessary to capture the accurate stress-strain behavior and

damage mechanisms of woven composites, they can hardly be used for analysis

of macroscale structures due to high computational cost. Thus, even for static

loading, these models are seldom used for larger length scale of the woven

composites.

In general, the FE modeling of impact on woven laminates is implemented at

macroscale using energy based damageable material laws such as stiffness

degradation, cohesive law, etc [157–160]. The woven fabric lamina in these

models are represented using homogenized brick or shell elements. These

models are able to capture the global elastic-plastic behavior as well as damage

on the plies of the woven composites in impact loading with good accuracy and

less computational time. For example, Iannucci [160] proposed a numerical

modeling methodology based on continuum damage mechanics to model non-

linear material behavior and progressive failure of woven carbon fiber

composites under impact. A material model for implementation into FE

simulation based on in-plane continuum damage mechanics for woven fabric


Chapter 2 57

composites was proposed by Johnson [158]. This model is a generalization of

the methods developed for UD ply composites and contains in-plane elastic

damage in fiber direction, with a non-linear elastic-plastic model for the shear

response, in which the fiber and shear damage modes are assumed to be

decoupled. The plasticity and shear damage parameters in this model are

determined experimentally. This approach is simple but effective that can model

a wide range of impact problems on woven composites when tearing and

penetration are major failure mechanism.

2.7. Summary
In the past decade, the use of natural fibers from plants as reinforcement in

polymer composites have become more popular thanks to their promising

properties such as high specific stiffness, good acoustic and vibration damping,

and eco-friendly characteristics. Amongst the commonly studied natural fibers,

flax fiber is recognized as one of the high-performance fibers in terms of

strength and stiffness per density, which makes it suitable for a number of

structural applications. In this regards, proper numerical models to predict the

mechanical properties of flax fiber composites can provide a better

understanding of their behavior and failure under different loading conditions.

The availability of these models can help to extend their use to engineering

applications.

To utilize the maximum load carrying capacity of the natural fibers, they must

be used in the form of continuous unidirectional (UD) or woven textile (with


58 A Literature review

optimally twisted yarns) in the composite laminates. However, one of the most

concerned failure modes of laminated composites is interlaminar failure

(namely delamination) since there is no reinforcement in the thickness direction.

Through-the-thickness stitching as a technique able to improve the performance

of composites has been around for some time. Stitching of the fiber preforms

has been acknowledged as a convenient and cost-effective technique for

improving the interlaminar properties, impact resistance and damage tolerance.

However, it is found that stitching may cause degradation in some mechanical

properties of composites such as tensile strength. As of today, a considerable

amount of experimental and numerical studies have been devoted to

investigating the influence of stitching on mechanical properties and impact

behavior of synthetic composite laminates. However, the effect of stitching on

the properties of natural fiber composites has been not comprehensively studied.
Chapter 3 59

Chapter 3
Experimental Methodology

3.1. Introduction
This chapter will provide an overview of the experiments and the numerical

methods used. The material, composite fabrication process, specimen

preparation and experiments, and the numerical modeling strategy for DCB

tests of the stitched composite are described in detail. Firstly, in-plane fiber

reinforcements, stitching materials and the thermoset resin system used are

introduced. Secondly, the detail of stitching procedure of fiber preforms,

stitching parameters and stitch areal fraction definition, and fabrication of

composites will be explained. Thirdly, the methodology of experiments and

measurements carried out in this study are described. Finally, the simulation

strategy of DCB test of the stitched flax fiber composite is presented. These are

inextricably linked, where the numerical modeling are developed based on the

experimental observations. It should be noted, however, that the detailed

description of the experimental results has been discussed in later chapters.


60 Experimental Methodology

3.2. Material
3.2.1. In-plane flax fiber reinforcements

In this study, flax fibers were used to reinforce the composites. Among the

commercial available long flax fibers for higher performance biocomposite

applications, Woven 4×4 hopsack flax fabric with an area density of 500g/m²

supplied by Composite Evolution (UK) [168] and 110g/m² unidirectional flax

fibers with a high degree of fiber alignment supplied by LINEO (Belgium) [169]

were employed as in-plane reinforcements, as shown in Figure 3.1. The woven

fabric uses a unique twistless yarn to provide optimum performance of the

natural fiber. For benchmarking purposes, unidirectional E-glass fibers of

498g/m² from Formax (UK) was also used to make the synthetic fiber composite.

(a) (b)

Figure 3.1. (a) Biotex flax 4x4 hopsack fabric; (b) FlaxTape©110 unidirectional fiber.

3.2.2. Stitch fibers

Two types of natural fiber rovings were used to stitch preforms (Figure 3.2): a

Tex 250 twistless flax yarn supplied by Composite Evolution (UK) [168], and

alternatively, Tex 30 twisted cotton thread produced by Gütermann (Greece)

[170]. The flax stitch yarn, which is the same yarn used to produce the woven
Chapter 3 61

hopsack flax fabric, is a bundle of technical fibers (dry roving) that are bound

by thin polyester thread. The cotton thread is dry spun, low cost, readily

available and was chosen as a stitch fiber to provide a lower bound reference

from the range of natural fibers available in the market.

(a) (b)

Figure 3.2. (a) Biotex 250tex flax yarn; (b) Natural cotton C Ne 50 thread.

3.2.3. Thermoset matrix

A resin-infusion grade thermo-set resin and hardener Epolam 5015 supplied by

Axson (France), with a mixed viscosity of 210 mPa.s (at 25℃) was used as the

polymer matrix. This resin is a low-viscosity system designed for liquid

composite molding and appropriate for wood and natural fibers [171]. The mix

ratio by weight was 10:3. The mixed resin must be cured at least 2 hrs at 50℃

or 24 hrs at room temperature to be demolded, and must be postcured (at least

16 hrs at 80℃) to obtain the optimal mechanical properties.


62 Experimental Methodology

3.3. Preform stitching


A range of unstitched and stitched laminates with various fiber architecture and

stitch densities were prepared in order to study the effects of stitch material and

density on the static and impact behavior of the flax fiber composite laminates.

To study the stitch fiber areal fraction effect, the stitching of woven flax

preforms was carried out at four different densities using Tex 250 flax yarn and

Tex 30 cotton thread.

3.3.1. Cotton thread stitching

The cotton thread stitching process was conducted using a commercial sewing

machine and universal needle. This method is appropriate for thin and flexible

un/twisted threads and is able to stitch laminates with medium thickness (up to

4-5mm). Figure 3.3 shows the stitching process using a sewing machine and

cross-sectional micrograph of the woven flax composite laminates stitched by

cotton thread.

(a) (b)

Cotton thread stitch

Figure 3.3. (a) Stitching using sewing machine; (b) Cross-section of the cotton-thread stitched
woven flax/epoxy composite.
Chapter 3 63

3.3.2. Flax yarn stitching

Similar to carbon fiber stitch thread, the commercial sewing machine with

universal needle mechanism (used for cotton thread stitching) was not able to

stitch using the twistless flax yarn due to friability and lower flexibility of the

yarn. Weinberg et al. [111,112] invented a sewing machine needle and a sewing

mechanism which is useful for stitching and joining fabrics with a brittle thread.

The invented open eye needle and the special sewing machine for stitching

brittle thread are diagrammatically depicted in Figure 2.16.

For preparing a few preforms to conduct mechanical tests, the flax yarn stitching

was performed manually by imitating this stitching process method. Figure 3.4

shows cross-sectional micrograph of the woven flax composite laminates

stitched by flax yarn.

Flax yarn stitch

Figure 3.4. Cross-section of the flax-yarn stitched woven flax/epoxy composite.

To enable the comparison between various stitch densities, the following

dimensionless formula was used to define stitch fiber areal fraction:


64 Experimental Methodology

2𝐴𝑠 2𝐴𝑠
Stitch Areal Fraction = | + |
𝑆𝐿 × 𝑆𝑅 Stitching in X direction 𝑆𝐿 × 𝑆𝑅 Stitching in Y direction
(3.1)

where 𝐴𝑠 is the average cross-section area of the stitch (values in Table 4.1),

and 𝑆𝑅 and 𝑆𝐿 are the stitch row spacing and stitch length, respectively

(Figure 3.5b-c). The modified lock stitch process was chosen to perform in order

to minimize the defects inside the laminate by placing the loops between the

bobbin and needle yarn at the outer surface of the laminate [27]. A schematic of

the modified lock stitched preform and stitch parameters are presented in

Figure 3.5. The cross-sectional micrographs of the woven flax composite

laminates stitched by flax yarn and cotton thread (as given in Figure 3.3b and

Figure 3.4b) showed good consistency and well impregnated stitched fibers.

(a)
Stitch lines

Z (Stitch )
Y (Weft ) X (Warp )
Chapter 3 65

(b)
Stitch length
(SL)
Needle thread

Weft fibers

Z (Stitch)
Warp fibers

X (Warp )
Bobbin thread

SL
SL

• • • •
Stitch row spacing
(SR) • • • • SR
• • • •
• • • •
Y (Weft )
0.5 As
X (Warp ) Stitch direction

Figure 3.5. (a) Schematic view of a stitched preform of woven flax fiber; (b) Definition of
stitch parameters shown from side and top view.

In addition to the woven laminates, a flax yarn stitched UD flax laminate at the

same stitch areal fraction (0.0138) as a stitched woven flax laminate was also

prepared to observe the effect arising from the choice of preform fiber

architecture. Table 3.1 shows the details of laminates manufactured for this

study.

It is worth mentioning, in preparing the preforms for the DCB test specimens,

a region was left unstitched to accommodate a polytetrafluoroethylene (PTFE)

film insert that acts as a crack starter. The entire preform was stitched for the

tensile test specimens.


66 Experimental Methodology

Table 3.1. Summary of the composite lay-up and stitch parameters of the manufactured
composite laminates.

Stitch Stitch row Stitch Areal


Lamina Laminate Stitch Stitch
Layup length spacing Fraction
preform nomenclature material direction
(mm) (mm) x10-2

PW0 [0]4 Unstitched - 0 0 0

PWC1 [0]4 Cotton X 3 8 0.46

PWC2 [0]4 Cotton X 3 4 0.92

PWC3 [0]4 Cotton X,Y* 3 6 1.22


Flax
4×4Plain
PWC4 [0]4 Cotton X,Y* 3 4 1.83
Weave
(500 g/m2)
PWF1 [0]4 Flax X 12 8 0.69

PWF2 [0]4 Flax X 8 8 1.03

PWF3 [0]4 Flax X 12 4 1.38

PWF4 [0]4 Flax X 4 8 2.06

[0]16 - 0 0 0
Flax UD
UD0 Unstitched
(110 g/m2)
[0/90]4s - 0 0 0

UDF [0]16 Flax X 12 4 1.38

Glass UD
UD glass [0]10 Unstitched - 0 0 0
(498 g/m2)
* Stitching conducted at both directions to achieve a similar stitch areal fraction due to smaller As of cotton
thread.

3.4. Composite manufacturing


The composite panels were made by the Vacuum Assisted Resin Infusion

(VARI) technique, as shown in Figure 3.6. All fiber preforms were dried at 80℃

for 3 hours in a vacuum oven to reduce fiber moisture content prior to the

infusion set up. The preforms were then placed on an aluminum plate, layered

on with peel ply and breather, and vacuum bagged for 2 hours at room
Chapter 3 67

temperature prior to infusion. The epoxy mixture was also vacuum degassed for

30 min at room temperature.

After resin infusion, the composite was cured at 25 ℃ for 24 hours, de-molded,

then post cured in a convection oven at 80℃ for 16 hours. Specimens were cut

from the panel using water-jet, and dried immediately at 60℃ for at least 24

hours to remove any surface moisture absorbed.

Vacuum Hose

Resin Feed

Figure 3.6. The vacuum assistant resin infusion setup used for manufacturing composite
laminates.

From the optical and scanning electron microscopic (SEM) images in Figure 3.7,

it was observed that this method of manufacturing produced well infused

technical fibers where almost all the single fibers are found to be wetted by the

matrix.
68 Experimental Methodology

(a) Flax
elementary fibres

Lumen

Flax
technical fibre

Matrix

(b)

Resin infused
technical fibre

Lumen

(c) Resin impregnation of


technical flax fibre

Figure 3.7. (a) and (b) Optical micrograph of the cross-section of the flax fiber composite; (c)
SEM image of a fractured resin-impregnated flax fiber in the composite.
Chapter 3 69

3.5. Fiber volume fraction measurement


The fiber volume fraction ( 𝑉𝑓 ) of the composites were determined from

measured values of their weight to 0.001g and density via Eq.(3.2) in which the

void content has been neglected. Gas pycnometer was used to measure the

average densities of pure cured matrix (𝜌𝑚 ), dry fiber (𝜌𝑓 ) and the composite

(𝜌𝑐 ). An average 𝑉𝑓 of 31±2% was obtained for the woven fiber composite and

40±2% for the unidirectional fiber composites. The stitch fibers were

considered in the determination of the 𝑉𝑓 , and their contribution was not more

than 9% of the 𝑉𝑓 for the case with highest stitch areal fraction (PWF4).

𝜌𝑐 = 𝑉𝑓 𝜌𝑓 + (1 − 𝑉𝑓 )𝜌𝑚
(3.2)

3.6. Composite tensile testing


The tensile properties of the composite laminates in the X (warp) direction were

evaluated by tensile test according to ASTM D3039. The tests were conducted

on a Shimadzu universal testing machine equipped with a load cell of 25 kN at

room temperature, as shown in Figure 3.8. The tensile tests were carried out at

a cross-head displacement rate of 2 mm/min, and strain gauges were used to

record the deformation in the both longitudinal and transverse directions. The

tensile stiffness was obtained from the slope of stress-strain curve between 0.1%

-0.3% strain. The test specimens with a dimension of 250mm by 25mm were

cut from unstitched and stitched composite panels with an average thickness of

4mm. A minimum of 5 specimens were tested for each sample to obtain a

reliable standard deviation.


70 Experimental Methodology

Back surface

Figure 3.8. Tensile test of the woven flax fiber composite.

3.7. Strength evaluation of the stitch fibers


The stitch cotton thread and flax yarn, as a through-the-thickness reinforcement

of the composite, will experience tensile loading during the Mode I opening test.

It was, therefore, necessary to evaluate the tensile properties of the stitch fibers

before and after impregnated with resin. This information is needed to

understand the stress-strain contribution of the stitch reinforcements when

bridging a crack. The impregnated test specimens were prepared by infusing 25

cm long dry cotton thread and flax yarns with the same epoxy resin, Epolam

5015, via the VARI technique. The cured, impregnated fibers were tested at a

gauge length of 10 mm and the cross-head speed of 1 mm/min, using an Instron

5500 micro tester machine with a 1 KN load cell. The tensile tests were

conducted on ten specimens to assess the variation of the tensile properties of


Chapter 3 71

the stitch fibers. The schematic of specimen mounting tab and the tensile test

set up for stitch fiber are shown in Figure 3.9.

(a) (b)
Resin impregnated
Griping area
flax yarn

Paper frame

Cutting line
after clamping
Adhesive

Figure 3.9. (a) Schematic view of the mounting tab and (b) the tensile test of stitch fiber
specimen.

The fracture energy of the impregnated stitch materials was estimated from the

load-displacement curve of the tensile test [122]. The absorbed energy 𝐸 is

calculated by integrating the area under the load-displacement curve using

Eq.(3.3)


𝐸 = ∫ 𝐹(𝛿)𝑑𝛿
0
(3.3)
where 𝐹(𝛿) is the tensile force as a function of displacement δ.

3.8. Flexural testing


The flexural properties of the composite laminates in the X (warp) direction

were measured by three-point bending tests according to ASTM D790 [172].

Specimens of 150mm by 20mm by 4 mm were tested with a span length of


72 Experimental Methodology

120mm using a Shimadzu universal testing machine and a load cell of 25 kN at

a loading rate of 5mm/min. In order to obtain a reliable standard deviation, a

minimum of 5 specimens were tested for each sample. The maximum flexural

stress sustained by the test specimen during bending was reported as flexural

strength. The flexural modulus (i.e. modulus of elasticity in bending [172]) was

calculated based on the slope of the secant line to the flexural stress-strain curve

between 0.3% -0.5% strain.

Figure 3.10. Three-point bending test of the woven flax fiber composite.

3.9. Composite double cantilever beam (DCB) testing


The Mode I interlaminar fracture toughness was determined from the DCB test

in accordance with ASTM D5528-01 [173]. The tests were conducted on an

Instron 4505 universal testing machine equipped with a 1 kN load cell, at the

test rate of 1 mm/min. In order to prevent large deformation and failure of the

specimen arms, unidirectional CFRP tabs with an average thickness of 2.3 mm

were bonded to both sides of the woven flax fiber DCB test specimens using the

Scotch-Weld™ DP-420 Epoxy adhesive. For UD flax fiber composite

specimens, three plies of the UD glass fiber were added to both faces of flax
Chapter 3 73

fiber preform prior to resin infusion process. The final thickness of the glass

fiber backings (after curing) was measured to be 1 mm.

The aluminum load blocks were bonded to both arms of the specimens using

the same adhesive. All bonding surfaces were lightly polished, sand blasted and

finally wiped with acetone to remove contamination. The schematic of CFRP

tabbed DCB test specimen with the nominal dimensions is illustrated in

Figure 3.11.

Figure 3.11. Schematic of tabbed double cantilever beam (DCB) test specimen.

To increase contrast for visual crack length determination, a thin layer of white

correction fluid was painted onto the side of the specimen, and marked with an

ultra-fine point marker at 1 mm intervals, taking reference from the load line.

The initial loading-unloading cycle was performed to propagate the crack by 3

to 5 mm from the pre-delamination insert so as to create a naturally sharp crack

tip. During the test, the crack propagation was monitored visually using a

traveling microscope and the crack length correlating to the load and
74 Experimental Methodology

displacement values was recorded for every 1 mm of crack length increment. A

second loading cycle was then performed for the crack propagation of about 60

mm. After testing, the specimens were fully fractured to assess the actual

position of the PTFE insert and to verify propagated crack length. The DCB test

setup and specimen are shown in Figure 3.12.

Travelling microscope

Figure 3.12. DCB test setup and mounted specimen.

The Mode I strain energy release rate, GI, was calculated from the Modified

Beam Theory (MBT) [173],

3𝑃𝛿
𝐺𝐼 =
2𝑏(𝑎 + |∆|)
(3.4)

where 𝑃 is the opening load, 𝛿 is the crack opening displacement at load point,

𝑏 is the specimen width, 𝑎 is the crack length, and ∆ is the intercept of the plot

of the cube root of the specimen compliance against the crack length.

Before testing, the specimens were placed in the oven for 24 hrs at 60℃ to dry

the moisture induced during the cutting process. All the tests were conducted at
Chapter 3 75

temperature of 23-25℃ and a relative humidity of 65%. The specimens were

exposed to test condition at least 48 hrs prior to testing.

3.10. Low-velocity impact testing


Drop-weight impact test was performed to investigate the behavior of the

unstitched and stitched woven flax/epoxy composites as well as cross-ply

flax/epoxy composites under low-velocity impact loads in accordance to ASTM

standard D7136 [81]. Three specimens were tested for each type of sample, at

each energy level. A homemade drop-weight impact test apparatus with a

double column guide dropping weight mechanism and a hemispherical head

steel impactor of 9.99 kg weight was used. The impactor was equipped with a

Kistler 9333A force sensor and had a striker head diameter of 20 mm.

The specimens were placed in the impact support fixture with a cut-out

rectangular hole of 75 mm by 125 mm and clamped using rubber tip toggle

clamps at four points with an identical clamp force, as shown in Figure 3.13. A

rebound brake was used to prevent unintentional multiple impacts on the

specimens. After testing, the residual depth of the depression formed by the

impactor was measured at impact location using depth gauge micrometer.


76 Experimental Methodology

Force Sensor

Impactor striker Toggle clamps with


rubber tip

Specimen

Figure 3.13. Drop-weight impact test setup with a mounted woven flax specimen.

The impact force-time history was obtained from the impactor force sensor and

recorded every 5×10-3 ms (200kHz) for 50 ms duration to collect enough data

point before and after the impact event. The contact velocity was measured

using high-speed camera with 15000 frames/second. The incident energy of

19.6J and 44.6J was chosen to produce non-perforation but visible damage, and

full perforation impact, for all samples respectively. The drop height (H) was

calculated to generate the desired energies using Eq.(3.5)

𝐻 = 𝐸𝑖𝑚𝑝𝑎𝑐𝑡 /𝑀𝑔
(3.5)

where 𝐸𝑖𝑚𝑝𝑎𝑐𝑡 is the impact energy, 𝑚 is the mass of impactor and 𝑔 is the

acceleration of gravity (9.81 𝑚/𝑠 2 ). The impactor displacement versus time,


Chapter 3 77

𝛿𝑖𝑚𝑝 (𝑡), during contact was generated using the numerical integration of force-

time history as in Eq.(3.6)

𝑡 𝑡𝑃
𝑖𝑚𝑝 (𝑡)
𝛿𝑖𝑚𝑝 (𝑡) = 𝑉𝑐 𝑡 + 0.5𝑔𝑡 2 − ∫ ∫ 𝑑𝑡𝑑𝑡
0 0 𝑚
(3.6)

where 𝑉𝑐 is contact velocity and 𝑃𝑖𝑚𝑝 (𝑡) is the incident force at time 𝑡. The

displacement obtained from Eq.(3.6), was validated with the data obtained from

the high speed camera. The absorbed impact energy was determined by

calculating the area under the force-displacement curve.

After impact testing, all the specimens were visually checked for any external

damage on both impacted and back surfaces. The damage type as well as the

length of cracks were recorded for all the specimens. A digital camera was used

to capture the scaled images of both surfaces of specimens before and after

impact.

3.11. Scanning Electron Microscopy


The fracture surface/morphology of the composites was analyzed by Scanning

Electron Microscopy (SEM). Using this technique, it is possible to inspect the

morphology of the flax fibers and their composites and identify the fracture

mechanisms which have a major contribution in energy dissipation.

The SEM samples were cut from different parts of the fracture surfaces of the

composites and then fixed by a double-side carbon tape to a specimen holder.

Prior to SEM imaging, an ultra-thin coating of gold (Au) was applied onto the

samples using sputter deposition technique for 100-120 sec at a vacuum


78 Experimental Methodology

pressure of 10 mbar and sputtering current of 15 mA in order to create a

conductive layer of metal to prevent charging during SEM. The SEM imaging

was performed using a Zeiss EVO extended pressure electron microscope in

secondary electrons mode which scans the sample surface with a high-energy

beam of electrons in a raster scan pattern. The images were taken at accelerating

voltage of 2-5 kV and working distance of 9-12 mm.


Chapter 4 79

Chapter 4
Mechanical and fracture properties of

stitched flax fiber composites

4.1. Introduction
This chapter presents the experimental work conducted to determine the

properties of the stitch fibers, and the mechanical properties of the stitched

composite, to study the impact of employing stitched preforms. These

experimental measurements also serve to provide the data required in the finite

element modeling and validation work presented in the next chapter. First, the

evaluated tensile properties of the stitch materials, before and after resin

impregnation are presented. This is followed by the tensile, flexural and Mode

I interlaminar fracture toughness properties of various stitched woven flax fiber

composite systems. The interesting effects of stitch areal fraction, the stitch

material selected and of the influence of in-plane preform architecture on the

fracture toughness and behavior will be discussed in detail. Finally, the results

are also benchmarked with the UD glass fiber composites to put the strength

and limitations of this natural fiber composite in perspective.


80 Mechanical and fracture properties of stitched flax fiber composites

4.2. Strength of dry and impregnated stitch fibers


The measured tensile properties of the twist-less flax yarn and the cotton thread

both in dry and impregnated form (composite) are shown in Table 4.1. This

information is useful for understanding the strength and fracture energy

contribution of through-the-thickness stitch reinforcements when bridging a

crack since the failure of all the stitches occurs under tension at the delamination

plane. More discussion on this will be made in next sections.

As it can be seen from Table 4.1, the strength and fracture energy of the

impregnated cotton threads are lower than those of the impregnated flax yarn.

Comparing the tensile properties of both stitch fibers before and after

impregnation, the strength and force at break of the impregnated flax yarn show

higher improvement than the impregnated cotton threads, over their dry

versions (Figure 4.1a). This discrepancy between the properties of flax and

cotton stitch fibers can be attributed to the particular form of the dry roving flax

yarn which is a bundle of unidirectional technical fibers bound by a thin thread,

as shown in Figure 4.1b. The length of these technical fibers normally varies

from 30 mm to 100mm. In a dry yarn, tensile load is taken up by the fibers

mainly through mechanical interlock and friction between the fibers, hence the

maximum load capacity of the fibers is not utilized. However, in a well-

impregnated yarn, the load is distributed and transferred among the fibers

through the matrix.


Chapter 4 81

(a)
120
Flax yarn (Dry)

100 Flax yarn (Impregnated)


Cotton thread (Dry)

80 Cotton thread (Impregnated)


Load (N)

60

40

20

0
0 2 4 6 8 10 12
Elongation (%)

(b) (c)

Binding thread

Figure 4.1. (a) Comparison of load-elongation curves of cotton thread and flax yarn before and
after resin impregnation; (b) dry twist-less flax yarn with binding thread; (c) dry twisted cotton
thread.

In the case of cotton thread, there is also an improvement in the strength of the

impregnated thread compared to that of the dry version, but the mechanism is
82 Mechanical and fracture properties of stitched flax fiber composites

different from that of flax yarns. The structure of cotton thread is made of spun

and twisted fluffy staple seed fibers with no axially-aligned technical fiber,

which explains the higher elongation of the thread before breaking because the

fibers can be straightened under load, Figure 4.1c. The strength of the dry thread

is high due to high friction between the twisted fibers which leads good load

transfer between the fibers. However, due to the high level of twist, the

permeability of the cotton thread is low and consequently its impregnation is

more difficult [6]. Hence, the improvement in the strength of the cotton thread

after impregnation by resin is not as significant as the untwisted flax yarns.

These results imply that natural fibers have to be used with the lowest level of

twist when they are to be impregnated with polymer resin as reinforcement. This

is unlike the textile applications where a high level of twist is necessary to allow

the dry fibers become usable on textile machines.

4.3. Effects of stitch areal fraction on the in-plane


properties of the woven composite
The effects of through-the-thickness stitch reinforcements on the in-plane

tensile and flexural properties of the stitched composite laminates were

investigated. For each sample at least five specimens were tested, and the

average values as well as the scatter of results (maximum-minimum values)

were presented. The samples comprised of flax yarn and cotton thread stitched

laminates.
Chapter 4

Table 4.1. Tensile properties of cotton thread and flax twist-less yarn.

Before Impregnation After Impregnation


linear density Cross-section area
Material
(Tex) (mm2) Elongation at break Force at break Strength Elongation at break Force at break Strength Fracture energy Vf
(%) (N) (MPa) (%) (N) (MPa) (Nmm) (%)
Flax Yarn 250 0.33 0.02 5 0.4 20 1.4 60 5.6 3.7 0.4 109 12.5 330 42 20.2 3.2 35

Cotton thread 30 0.055 0.01 11 0.5 8 0.4 145 27 10 0.8 13 1.5 230 48 6.5 0.9 45
83
84 Mechanical and fracture properties of stitched flax fiber composites

4.3.1. Tensile properties

Figure 4.2a and b illustrate the tensile strength and stiffness versus the stitch

areal fraction of both cotton and flax stitched composites, respectively. The

graphs indicate that the tensile properties, regardless of the types of stitch fibers,

are expected to decrease gradually with increasing density of stitching (stitch

areal fraction). Only the cotton thread stitched composite of lowest stitch areal

fraction (0.0046) showed lower strength than both the unstitched and the next

higher stitch areal fraction (0.0092). This variation is likely to have been

influenced by the manufactured quality – i.e. void content and fiber

misalignment – associated with the individual panels, as the composite panels

of each stitch areal fraction have been processed separately. The detailed tensile

test results are reported in Table B.1 in Appendix B.

(a)
95

90

85
Tensile Strength (MPa)

80

75

70

65

60
Cotton thread stitch
55
Flax yarn stitch
50
0.000 0.005 0.010 0.015 0.020 0.025

Stitch areal fraction


Chapter 4 85

(b)
9

8.5

8
Tensile Stiffness (GPa)

7.5

6.5

5.5 Cotton thread stitch


Flax yarn stitch
5
0.000 0.005 0.010 0.015 0.020 0.025

Stitch areal fraction

Figure 4.2. Effects of various stitch areal fractions of cotton thread and flax yarn on the tensile
properties of woven composite; (a) tensile strength; (b) tensile stiffness.

The results for tensile stiffness showed no change or even a modest

improvement up to stitch areal fractions of 0.01 for both stitch fibers. For

example, the tensile stiffness of the stitched samples PWC2 and PWF2 (with a

similar stitch areal fraction of 0.01) were slightly improved against the minor

reduction (up to 4%) in the strength. A similar trend was also reported for the

stitched woven synthetic fiber laminates in [105], in which the mechanical

properties of the stitched woven composite laminates were improved at an

optimum stitch areal fraction compared to the unstitched one. The plausible

mechanism for the modest improvement in the in-plane stiffness of stitched

composites is the increase in the in-plane fiber volume fraction due to

compaction of the fiber preforms by the tension in the stitch fibers during

processing [32] (see Chapter 2.5.1.2).


86 Mechanical and fracture properties of stitched flax fiber composites

It is apparent from Figure 4.2b that the tensile stiffness of the two types of

stitched laminates exhibits similar levels of decline with increasing stitch areal

fractions from 0.01. This is despite the fact that the flax yarn size is six-fold

bigger than cotton thread. In fact, for cotton thread, due to the lower cross-

sectional area and higher stitch punch frequency, the physical displacement of

the in-plane fibers is less per stitch but more numerous sites. Conversely, in flax

yarn stitching, the distortion of in-plane fibers is more per stitch because of its

larger diameter, but fewer in numbers of stitch. The overall effect may be

compensated so that the difference between the two cases becomes less

pronounced.

The results in this study showed that the maximum reduction in tensile

properties for the both types of stitch occurs at the highest density of stitching.

The observed reduction in the tensile strength is not as severe as the results

reported for some of the stitched synthetic fiber composite laminates (see

Figure 2.21) [106,123]. Reviewing the available data on the conventional

stitched composites, the in-plane mechanical properties of stitched composites,

depending on the stitching parameters (i.e. stitch fiber material, dimension and

stitch areal fraction) and the in-plane fiber architecture, are generally degraded

after stitching [21].

Generally, the reduction in the tensile properties of stitched laminates is

attributed to geometrical stitch-induced defects such as crimping and distortion

(or waviness) of in-plane fibers, and in-plane reinforcing fiber breakage during

the stitching process [25,32]. Nevertheless, it was found that the reduction in

the tensile properties of both types of the stitched composites was very similar.
Chapter 4 87

This observation suggested that the in-plane fiber damages of the woven flax

are not very susceptible to the punching frequency of the sewing needle through

the preform as a higher number of cotton stitches per unit area (i.e. more needle

damage) to achieve a similar stitch fiber areal fraction to the flax stitched

laminate.

This behavior of the woven flax preform may be due to the inherent variation

in the length and cell wall thickness of the flax fiber bundles [44] as well as

reasonable slack in the fiber-bundles of this particular 4x4 hopsack fabric weave

which allowed the elementary fibers to move out of the way as the needle

punches through. Hence, placing a stitch thread or yarn through it would have

caused minimal distortion and breakage to this particular weave type of in-plane

fibers.

However, in-plane fiber crimping and distortion are inevitable consequences of

stitching that would have led to the formation of resin rich areas and stress

concentrations around the stitch fibers. In addition, stitching also made the

stacked fiber layers more compact and this inversely affected the resin

permeability through the fiber preform during fabrication. These factors would

have increased the presence of process defects such as voids, porosity, and

matrix micro-cracking in the vicinity of the stitch fibers in the composite

[32,119] with increasing stitch areal fraction and thereby led to the

corresponding decline in the tensile properties.


88 Mechanical and fracture properties of stitched flax fiber composites

4.3.2. Flexural properties

The flexural test results of unstitched and stitched woven flax fiber composites

are given in Figure 4.3a and b. The detailed flexural test results are given in

Table B.2 in Appendix B. The flexural strength of the cotton thread stitched

laminates showed almost no change against the different stitch areal fraction,

while a decline in the flexural strength of up to 12% (at the highest stitch areal

fraction) was observed in results of the flax yarn stitched laminates. The flexural

modulus of the both types of stitched laminates remained unchanged with

increasing stitch areal fraction.

(a)
140

130
Flexural Strength (MPa)

120

110

100

90 Cotton thread stitch


Flax yarn stitch
80
0.000 0.005 0.010 0.015 0.020 0.025

Stitch areal fraction


Chapter 4 89

(b)
7

6.5

6
Flexural Modulus (GPa)

5.5

4.5

Cotton thread stitch


3.5
Flax yarn stitch
3
0.000 0.005 0.010 0.015 0.020 0.025

Stitch areal fraction

Figure 4.3. Effects of various stitch areal fractions of cotton thread and flax yarn on the
flexural properties of the woven composite; (a) flexural strength; (b) flexural stiffness.

It was observed that all fractures initiated at the bottom face of specimens which

is under tension in bending. The fiber failure under tension was the dominant

failure mode of the specimens, and there were no signs of compressive fiber

failure observed at the top plies of specimens. This is attributable to the high

crimp of woven preform fibers in this particular woven (4x4 hopsack) flax

fabric. The large deformation of composite ply due to buckling of fibers under

compression loading controls by the plasticity of matrix which is the dominant

failure mode at top plies in bending stress. The schematic of fracture mechanism

under three-point bending test is shown in Figure 4.4.


90 Mechanical and fracture properties of stitched flax fiber composites

Resin-rich area Matrix plasticity (Compression)

Fiber failure (Tension)

Figure 4.4. Schematic of three-point bending test of woven flax composite.

Unlike the tensile properties, the flexural strength of stitched laminates is more

sensitive to the thickness of the stitch fibers. The maximum difference between

the average of the flexural strength of cotton-thread and flax-yarn stitched

composites (at the highest stitch areal fraction) is 11%, while the same

comparison for the tensile properties results in a difference less than 5%.

The existing difference between the flexural strength of the flax yarn and cotton

stitched laminates can be associated with the dimension of the stitches. It was

observed from the analysis of fractured specimens that the failure under bending

initiated at the stitch loop of flax yarn at bottom ply and propagated then along

with the stitch direction (Figure 4.5a). This fracture mechanism was not

generalizable to the cotton thread stitched specimens as the crack did not occur

at the location of the cotton stitches (Figure 4.5b). This could be due to the larger

resin rich areas generally created around the flax yarn stitch loops close to
Chapter 4 91

specimen surfaces which leads higher stress concentration at that point [105].

During three-point bending, the maximum stress happens at the outer surfaces

of the center of the specimen and the influence of stress concertation, therefore,

is more significant. In the case of cotton thread stitching, the resin rich areas

around the stitch loops are minor due to the smaller diameter of stitch thread,

and consequently, the stress concentration is lower.

(a) Flax yarn stitch (b) Cotton thread stitch

Figure 4.5. Failure of (a) Flax yarn and (b) Cotton thread stitched woven flax composite under
bending stress.

4.4. Effect of stitch areal fraction on the Mode I


interlaminar fracture toughness of flax woven fabric
composite
The interlaminar fracture toughness of the stitched woven flax fiber composite

laminates was investigated through the DCB test described in Chapter 3.9, at

various stitch areal fractions for flax yarn and cotton thread stitched laminates.

The detailed DCB test results including the measured values of initiation and

propagation Mode I interlaminar fracture toughness, GIC, are reported in

Appendix B.
92 Mechanical and fracture properties of stitched flax fiber composites

Figure 4.6 shows the average values of the Mode I interlaminar fracture

toughness, GIC, obtained for both stitch fiber types, plotted against their stitch

areal fractions. Through-the-thickness stitching with cotton thread did not make

much enhancement to the GIC values as compared to the unstitched reference,

and the highest improvement attained was only about 4%. Unlike the cotton

thread stitched laminates, the flax yarn stitched laminates show a gradual and

continuous improvement of delamination resistance at increasing stitch areal

fraction, and achieved an increase of 21% at the highest stitch areal fraction

investigated.

It was observed during the DCB tests that the crack propagation of the cotton

stitched laminates was stable. Stable crack growth was also observed flax-

stitched laminates with lower stitch areal fractions (PWF1 and PWF2), but the

higher stitch areal fraction laminates (PWF3 and PWF4) exhibited stick-slip

crack propagation (unstable crack growth followed by crack arrest). When

analyzing stick-slip specimens, only the initiation and propagation data were

used to calculate the values of GIC (as per ASTM D5528 standard [173])
Chapter 4 93

4500
PWF4

PWF3
4000 PWF2
PWF1
GIc (J/m2)

3500
PW0

3000 PWC4
PWC2 PWC3
PWC1

2500
Cotton thread stitch
Flax yarn stitch
2000
0.000 0.005 0.010 0.015 0.020 0.025

Stitch areal fraction

Figure 4.6. Effects of stitch areal fraction and stitch material on the interlaminar fracture
toughness (G IC) of woven flax/epoxy composites.

The values of GIC versus crack length (R-curve) of the highest stitch areal

fraction of the cotton thread and flax yarn stitched laminate (PWC4 and PWF4)

were compared to the unstitched laminate in Figure 4.7. A number of

explanations can be offered for the discrepancy between the performance of the

flax yarn and cotton thread stitched laminates. First, from the tensile test results

of wet rovings in Table 4.1, the tensile loads supported by the flax yarn during

crack bridging is likely to be very much higher than that of the cotton thread.

The lower absorbed energy at the failure of the cotton thread compared to that

of the flax yarn (Table 4.1), may have rendered the cotton threads ineffective at

the breaking loads experienced by the tests. This could explain why the fracture

energies of the cotton stitched laminate (highest stitch areal fraction, PWC4) in

Figure 4.7 was not much different from the unstitched specimen (PW0), but the

flax yarn stitched laminate (PWF4) performed so much better.


94 Mechanical and fracture properties of stitched flax fiber composites

6000

5000

4000
GIc (J/m2)

3000

2000

1000
PW0 PWC4 PWF4

0
55 60 65 70 75 80 85 90 95 100 105

Delamination Length (mm)

Figure 4.7. R-curves of the stitched flax woven laminates using cotton thread and flax yarn at
almost equal stitch areal fraction of 0.02.

Second, the deteriorative effects of cotton stitching happened to cancel out any

interlaminar toughness improvements contributed by the bridging traction of the

cotton threads. As discussed previously, stitching would raise the incidence of

defects around the stitches (i.e. resin pockets, micro-voids and matrix micro-

cracks) within the composite by distorting and crimping the in-plane fibers as

well as reducing the resin permeability of the preform. In the presence of a

higher number of stitch punches per unit area, the higher concentration of defect

sites in close proximity in the cotton-thread stitched composites (Figure 4.8)

would have made it easier for the interlaminar micro-cracks to coalescence and

grow. Since the interlaminar toughness of the cotton-stitched laminates in

Figure 4.6 did not decrease but remained high and unaffected by the density of

stitching, a net toughening effect may be present, but due to low strength of

cotton thread, it was simply too small to show up statistically.


Chapter 4 95

(a) Delamination front


(b)

stitch-induced
defect

Interlaminar
micro-cracks

Figure 4.8. Comparison of the frequency of the stitch-induced defects at a similar stitch areal
fraction of (a) cotton thread and (b) flax yarn stitched composites.

Third, the bridging traction of the through-the-thickness flax yarn stitch fibers

is higher than those of the in-plane fibers but not for the cotton-thread stitch

fibers. In all the test specimens, a large-scale in-plane fiber bridging zone which

grows into a region several centimeters long at the wake of the crack front, was

typically observed during crack propagation. This extensive fiber bridging

activity encourages matrix shear yielding and significantly increases the energy

required to advance delamination. It normally supports stable crack propagation

through the flax composites but for the flax yarn stitched laminates, stick-slip

crack growth with crack initiation loads higher than the average stable

propagation loads (Figure 11) was a constant occurrence at stitch areal fractions

above 0.01. The thicker stitches (flax yarn) offering a combination of larger

fiber surface bonding area, higher breaking loads, intra-yarn fracture and slip-

bridging, are thought to contribute favorably to this successful stitch-toughening

outcome with the flax yarn.

The gain in Mode I fracture toughness through (flax yarn) stitching came at the

expense of the deterioration of some in-plane properties, hence there is a limit


96 Mechanical and fracture properties of stitched flax fiber composites

to the efficiency of this toughening method. Comparing the tensile (Figure 4.2a-

b), flexural (Figure 4.3a-b) and DCB test results (Figure 4.6), the highest rate of

increase of the GIC values occurred below the stitch areal fraction of 0.01. The

percentage gain in GIC was 16% from 0 to 0.01 (stitch areal fraction), while it

was only 4% from 0.01 to 0.02. Coincidentally, the rate of decline in the tensile

and flexural properties of the composites was lower at stitch areal fractions

below 0.01. Therefore, increasing the stitch areal fraction beyond 0.01 for this

composite system is not likely to be very meaningful. Preservation of the in-

plane mechanical properties of natural fiber composites, which are not high to

begin with, certainly favors lower stitch areal fractions. The change in failure

propagation mode from stable to unstable stick-slip crack growth above the

stitch areal fraction of 0.01 also discourages the use of very high stitch densities

to achieve higher interlaminar toughness. Moderate use of stitching to enhance

the delamination resistance at localized areas, may possibly be the best way to

employ the benefit of this approach.

4.5. Fracture mechanisms of Mode I DCB samples from


SEM micrographs
In Figure 4.9, a fractured cotton and flax stitch are shown in the SEM images of

delaminated surfaces of stitched woven flax fiber laminate. It could be observed

from the fracture surfaces that unlike the synthetic fiber laminates stitched by

strong materials (for instance Kevlar) that tend to break at the surface stitch loop

and be pulled out from the laminate (see Figure 2.24) [103,115], both the cotton

thread and flax yarn tend to break at the delamination plane, with the average
Chapter 4 97

fiber pull-out length not longer than a single ply thickness of in-plane fibers

reinforcing the substrate. The crack development within the flax yarn stitched

laminates is therefore likely to be similar to the steps that have been proposed

by Ogo [174], as shown in Figure 4.10.

(a) (b)

Fractured flax
Fractured cotton stitch yarn
stitch thread

Figure 4.9. SEM image of fractured stitches within the stitched woven composite; (a) cotton
thread, (b) flax yarn.

First, the delamination (matrix) crack passed round the stitch fiber and then was

arrested behind the stitch line. Then the stitch fiber stretched as it began to carry

some load and small debonding would occur along the fiber-matrix interface.

Some fiber yarn splitting and slippage was likely to occur at this point as well,

as seen from the sub-bundle fractures occurring along then stitch yarn in the

SEM image (Figure 4.9b). Strain energy would then be built up in the fibers and

the surrounding matrix. Finally, the stitch fibers fractured and released the

stored energy causing the delamination crack to propagate at higher speeds

towards the vicinity of the next stitch line, giving rise to the unstable stick-slip

crack growth behavior. Similar fracture mechanism is proposed for the cotton
98 Mechanical and fracture properties of stitched flax fiber composites

stitched laminates as well, but with much lower levels of strain energy build up

and release as a result of lower strength of the cotton thread.

A matrix crack passed round the The thread was stretched and The stitch thread failed and
stitch thread and then was some small debonding occurred the stored energy was released
arrested behind the thread. along the thread. to propagate the delamination
at higher speed until the next
stitch thread.

Figure 4.10. The 3-step delamination growth in the stitched flax fiber composites.

4.6. Effect of in-plane fiber architecture on the Mode I


interlaminar fracture toughness of unstitched and
stitched composite laminates
The influences of in-plane fiber architecture of stitched and unstitched

composite laminates on the Mode I interlaminar fracture toughness are

discussed in this section. Figure 4.11 shows some representative load-

displacement curves of both the unstitched and stitched laminates of UD and

woven flax fiber composites. Flax yarn was used to stitch both preforms with a

similar stitch areal fraction of 0.014. It can be seen that both stitched laminates

attained higher peak loads compared to their unstitched counterparts. Rapid

crack growth followed by crack arrest was observed during DCB test for

stitched laminates at this stitch areal fraction. However, the arrest loads never

dropped below the steady-state propagation loads of the unstitched laminates.


Chapter 4 99

300

PWF3 PW0 UDF UD0


250

200
Load (N)

150

100

50

0
0 5 10 15 20 25 30
Cross-head Displacement (mm)

Figure 4.11. Typical Mode I load-displacement curves of unstitched and stitched flax UD and
woven laminates.

The measured Mode I interlaminar fracture toughness against crack length (R-

curve) were plotted in Figure 4.12 to compare laminates with different in-plane

fiber architectures. Again, both the UD and woven flax laminates were stitched

with flax yarns to the same stitch fraction of 0.014. To compare the interlaminar

properties of the flax fiber composites with synthetic glass fiber composites,

glass fiber DCB composite specimens with the lay-up of [0]10 and 𝑉𝑓 of 61%

were also manufactured (using the same resin system and processing method)

and tested. This information is very helpful in understanding the influences of

in-plane fiber architecture as well as through-the-thickness stitching on the

delamination resistance of the flax fiber composites.


100 Mechanical and fracture properties of stitched flax fiber composites

5000

4500

4000

3500

3000
GIc (J/m2)

2500
PWF3
2000
PW0
1500 UDF

1000 UD0

500 UD glass

0
50 60 70 80 90 100 110 120
Crack Length (mm)

Figure 4.12. Comparison of R-curves of the flax and glass fiber composites.

It is clearly seen in Figure 4.12 that the UD glass fiber composites (denoted by

‘UD glass’) have significantly lower interlaminar toughness compared to all the

UD and woven flax fiber composites. The GIC values of UD glass fiber

composites are less than half of that of the UD flax fiber composites. This is

attributable to the large difference in the degree of fiber bridging across the

delamination plane of the flax and glass fiber composites. The high strength

glass fibers tend to remain unbroken, and the occasional dislodged chords are

also less mobile when bridging the crack. UD flax fibers on the other hand are

formed by shorter technical fibers with irregular length, diameter, local

alignment, and nests more deeply between plies. Their irregular geometry

creates uneven fracture surfaces and the broken flax fibers are able to pull out

and adjusts its bridging angle across the crack plane more easily. Figure 4.13

compares the fiber bridging observed during Mode I delamination of the woven
Chapter 4 101

and UD flax composites with the UD glass fiber composite, where the large

fiber bridging zones in the natural fiber composites can extend up to 30mm in

the wake of the crack tip.

Comparing the results of the interlaminar fracture toughness for the UD flax

and woven flax composites in Figure 4.12, it is apparent that the form of the

fabric makes a significant difference. The average GIC values of the flax woven

composite are about three times higher than that of the flax UD composite. The

main toughening mechanism appears to be the nesting phenomena, which is

more significant for the woven composites due to the waviness of the fabric

surface. Generally nesting as well as the presence of 90°-bundles in a weave

result a rugged crack path, consequently increases the energy required to

propagate crack in-between the plies [175]. The interlaminar fracture toughness

therefore rises when nesting occurs within the laminate. A high level of nesting

was observed in delamination of woven flax fiber composite compared to UD

configuration. This can be clearly seen from the delamination propagation path

of the studied composites in Figure 4.13. Due to this, extensive in-plane fiber

fractures were observed on the delaminated surfaces of the woven DCB test

specimens (Figure 4.14a-c), and therefore a substantial amount of energy would

have been absorbed by the fiber pull out, fiber breakage and fiber-matrix

debonding mechanisms during delamination. The large-scale fiber bridging

observed for the woven composites (Figure 4.13a) also continues to be the other

toughening mechanisms for the flax woven composites [35]. The different level

of fiber bridging observed in woven and UD composites could partially explain

the large difference between their interlaminar fracture toughness.


102 Mechanical and fracture properties of stitched flax fiber composites

(a)

(b)

(c)

Figure 4.13. Fiber bridging during delamination growth; (a) woven flax fabric, (b)
unidirectional flax, and (c) unidirectional glass fiber laminate composite.

Finally, an examination of the spaces between the warp and weft tows of the

woven flax fabric shows that they are being filled by resin and forming

relatively large resin pockets (Figure 4.14c). During delamination, these

interlayer resin pockets can be stretched by the surrounding fibers they are

bonded to and absorb significant amounts of energy through matrix shear

yielding (Figure 4.14d) [176]. Higher interlaminar fracture toughness of the

woven fabric laminates over non-woven UD lay-ups was also observed in

conventional composites, and the increase had been attributed to the inherent

roughness of the woven fabric and the larger plastic yield zone at the crack tip

due to a larger resin-rich region between plies [177,178]. The confluence of


Chapter 4 103

these energy intensive fracture mechanisms, particularly the presence of resin

pockets and larger plastic yield zone size, made the woven fabric architecture

more favorable for producing high toughness natural fiber composites.

(a) (b)

Fibre breakage

Fibre pull-out Fiber-matrix debonding

(c) (d)
Matrix shear yielding

Weft fibres

Resin

Warp fibres

Figure 4.14. SEM of a fractured surface of the flax woven fabric composite after DCB test; (a)
fiber pull out and breakage, (b) Fiber-matrix debonding,(c) a resin-rich region between warp
and weft fibers, (d) matrix shear yielding.

4.7. Conclusion
The effect of introducing out-of-plane natural fiber stitches on the Mode I

interlaminar fracture toughness as well as tensile and flexural properties of the

continuous flax/epoxy composite laminates were presented. The experimental

findings showed that the presence of stitches led to notable reductions in the in-

plane tensile strength and stiffness of the composites by as much as 18% and

12% respectively, regardless of the type of stitch material used. The flexural
104 Mechanical and fracture properties of stitched flax fiber composites

modulus of stitched composite remained almost unchanged. Nevertheless, the

flexural strength of the composite was reduced up to 12% when stitched by flax

yarn, while insignificant effect was observed for when stitched by cotton thread.

Stitching with the thicker 0.33 mm2 diameter flax yarn was found more effective

than the 0.055mm2 cotton threads in enhancing the Mode I fracture resistance

of woven flax/epoxy composites at similar stitch areal fractions. The Mode I

crack propagated in a stable progressive manner through cotton-thread-stitched

laminate at very similar fracture energies as the unstitched laminates. The flax-

yarn-stitched laminates, however, experienced a change in crack growth

behavior from stable propagation to stick-slip unstable fracture and the average

fracture loads measured (including the reduced loads at arrest points) were

higher than the average propagation fracture loads of the unstitched laminates.

This was observed in all the flax-yarn stitched composites having either

unidirectional or woven in-plane flax reinforcements. The ability to increase the

fracture energies by 21% was attributed to the combined factors of higher stitch

areal fraction, higher tensile loads that impregnated flax yarns (stitches) could

tolerate during crack bridging, and the presence of stitching-induced resin-rich

areas that appeared to support more extensive matrix shear yielding. One other

useful observation for the design of natural fiber composites with high Mode I

delamination resistance is that the woven preforms tend to promote greater

energy dissipation than the unidirectional flax preforms due to different

toughening mechanisms like higher level of nesting and fiber bridging.


Chapter 5 105

Chapter 5
Numerical modeling of Mode I interlaminar

failure of stitched flax fiber composites

5.1. Introduction
This chapter aims to propose a numerical approach to model the interlaminar

failure of stitched flax fiber composite laminates by simulation of the DCB test

studied in the previous chapter. The proposed approach to model the

interlaminar interface is constructed in two steps. In the first step, the

interlaminar failure of unstitched flax fiber laminate, as the parent laminate, is

modeled using cohesive elements with a nonlinear softening law in order to

model the large scale fiber bridging observed during the experimental study.

The experimental results are used to calibrate the parameters of the cohesive

law. In the second step, 2-node beam elements are superposed onto the cohesive

interface of the parent laminate at a prescribed distribution and density to model

the bridging stitches present in the validation samples. The stitch material

behavior and properties are obtained from the tensile test of impregnated stitch

fibers. The modeling strategy and material model used in this study are

described in Section 5.2. The model is evaluated based on the DCB test results

of the UD flax fiber/epoxy composite laminates. Finally, a parametric study is


106 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

performed on the tensile strength of the stitches in order to assess the effect of

variation in the strength of stitches on the interlaminar fracture toughness of the

composite.

5.2. DCB test simulation of stitched flax fiber composite


5.2.1. Modeling strategy

A DCB test specimen model with the fixed dimension of 6mm thickness, 20mm

width and 62mm initial crack length was created and analyzed using the

commercial software Abaqus/Implicit. A schematic illustration of the stitched

DCB test specimen and the stitch configuration used in the FE modeling are

shown in Figure 5.1.

Top view
Pre-crack area Stitch line
12 mm

90° Crack propagation direction 5×4 mm 20 mm


62 mm

170 mm

Side view Glass fiber Tabs [0]3

90°

Pre-crack area Needle yarn

Bobbin yarn

Figure 5.1. Schematic of DCB specimen of the stitched UD flax fiber composite, and stitching
parameters.
Chapter 5 107

The 3D finite element (FE) model was based on the experimental test specimen

of the UD flax fiber composite laminates backed with the UD glass fiber

composite tabs studied in Chapter 4. The 3D FE model used to analyze

delamination in stitched flax fiber composite is shown in Figure 5.2.

D
A
B E
C

LEGEND: A Cohesive layer B Flax fiber laminate C GFRP Tab


D Stitch fiber E Aluminum load block

Figure 5.2. 3D finite element model of the DCB test specimen. A finer mesh was used at the
stitched region.

To model the interlaminar failure behavior of the stitched flax fiber composite,

three fracture mechanisms associated with delamination propagation are

considered:

I. Decohesion of interlaminar interface (resin-rich layer between the fiber

plies) at delamination front.


108 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

II. Bridging traction of in-plane fibers in the wake of the crack

(delamination) tip.

III. Bridging traction of stitch yarns across the delamination plane.

II. In-plane fiber bridging III. Trough-the-thickness stitch I. Decohesion process

Delamination front

lpc2 lpc1 lpc1: Decohesion process zone


lpc2: Fiber bridging process zone

Figure 5.3. Three fracture mechanisms involved in delamination propagation.

Figure 5.3 illustrates the schematic of the fracture mechanisms during

delamination propagation. The mechanisms I and II are modeled using trilinear

cohesive law, and mechanism III is modeled by introducing beam elements.

The implementation of the model follows two steps. First, the trilinear cohesive

law is calibrated using the experimental results of the unstitched UD flax fiber

laminates (see Section 5.3.2). Then, 2-node beam element is superposed with
Chapter 5 109

the cohesive element to model the interlaminar interface of the stitched UD flax

fiber laminate.

In the FEM model, each arm of the specimen consists of two 8-node

quadrilateral in-plane continuum shell elements (SC8R) in the thickness

direction to model the UD composites of flax and glass fibers ([0Glass


3 /0Flax
8 ]s ), as

illustrated in Figure 5.2. The mid-section interlaminar layer where the two arms

are joined is modeled by an 8-node three-dimensional cohesive element

(COH3D8) with a very small thickness of 0.01mm. The cohesive elements are

connected to neighboring continuum shell elements by sharing nodes.

A trilinear cohesive law is used to model the effect of the fiber bridging during

delamination. 2-node beam elements (B31) are placed between the top and

bottom nodes of the cohesive layer to introduce through thickness stitch

reinforcements, following the distribution shown in Figure 5.1.

The boundary conditions are introduced by applying a vertical displacement on

the center nodes of the top load block and constraining the center nodes of the

bottom load block along the vertical direction. These boundary conditions are

according to the DCB experiment where the bottom load block of the specimen

is pinned to the fixed jaw, and the top load block is pulled using displacement

control.

An element size of 0.25 mm by 0.5 was used at the stitched region, and a coarse

mesh of 1 mm by 0.5 was used for the rest of specimen. A matched mesh was

used for both the continuum element and cohesive element. The element size

was determined based on the rule proposed by Harper [44] for the cohesive zone
110 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

length to ensure that the element is small enough to capture the stress gradient

at the wake of crack tips.

5.2.2. Material Model

5.2.2.1. Elastic behavior of laminate

The in-plane flax and glass fiber reinforced laminates of the arms were modeled

as a homogeneous orthotropic elastic material using Abaqus continuum shell

element. Each arm of the double cantilever beam specimen is modeled with

continuum shell elements. This model is accurate enough to capture the elastic

deformation of the arms of DCB test specimen [179]. The values of elastic

constants used in the modeling of flax and glass composite laminates are

summarized in Table 5.1.

5.2.2.2. Delamination

In this study, delamination is modeled using Abaqus cohesive element

(COH3D8). A trilinear traction-separation law implemented in a UMAT

subroutine is used as the cohesive law. The trilinear cohesive law is used in

order to model the R-curve effect of interlaminar fracture toughness of flax fiber

composite. This R-curve effect was due to large-scale in-plane fiber bridging

during delamination propagation, as will be discussed in Chapter 4 and 5 [180].

In the presence of an R-curve, the toughness measured during crack propagation

increases until reaching a steady-state value. Such response is generally because

of involving more than one physical phenomenon in the fracture process, some

acting at small opening displacements and others acting at higher opening

displacements and extending further into the crack wake. It has been shown that
Chapter 5 111

the use of a traction-separation law with a nonlinear softening law is necessary

to model the R-curve effect [150–152,155].

In this study, the approach proposed by Airoldi et al. [151] is applied to the

experimental R-curve to determine the parameters of the trilinear cohesive law

(𝑛 and 𝑚). This trilinear law is obtained from superposition of two bilinear

cohesive laws, as shown in Figure 5.4. The calibration of the superposition

parameters ( 𝑛 and 𝑚 ) for the UD flax fiber laminate is described in

Section 5.3.2.

𝜎𝑐

n𝜎 𝑐 = +

=
(1 − 𝑛)𝜎 𝑐 =( − )

𝛿𝑖𝑛𝑖. 𝛿𝑐1 𝛿𝑚𝑎𝑥

Figure 5.4. Superposition of two bilinear cohesive law and resultant trilinear cohesive law
(reproduce from [151]).

The cohesive tractions (𝜎𝑛 , 𝜎𝑠 , 𝜎𝑡 ) are related to the corresponding separations

in the normal and (both) shear directions (𝛿𝑛 , 𝛿𝑠 , 𝛿𝑡 ) thru Eq.(5.1), [179]

〈𝛿𝑛 〉
𝜎𝑛 𝐷𝑛𝑛 (1 − 𝑑) 0 0 𝛿𝑛
𝜎 |𝛿𝑛 |
[ 𝑠] = [ 𝛿𝑠 ]
𝜎𝑡 0 𝐷𝑠𝑠 (1 − 𝑑) 0
𝛿𝑡
[ 0 0 𝐷𝑡𝑡 1 − 𝑑)]
(5.1)
112 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

in which 𝐷𝑛𝑛 , 𝐷𝑠𝑠 and 𝐷𝑡𝑡 are the penalty stiffness, and 𝑑 is the damage

variable (𝑑 = 0 for when there is no damage in the interface and 𝑑 = 1 once

the interface is completely fractured). The 〈 〉 in Eq.(5.1) is Macauley bracket

operator which is defined for every 𝛼 ∈ ℝ as

(𝛼 + |𝛼|)
〈𝛼〉 = (5.2)
2

In other words the damage variable does not influence the 𝜎𝑛 when crack is

closed in compression.

The penalty stiffness 𝐷𝑛𝑛 , 𝐷𝑠𝑠 and 𝐷𝑡𝑡 are artificial parameters used to

constrain the separation (or interpenetration) between the crack faces and have

perfect-bonding between plies before delamination onset. Therefore, they must

be large enough compared to the stiffness of the lamina [181–183]. Here, the

formulas’ of Eq.(5.3) are used to calculate the minimum values of the penalty

stiffness [182]:

𝑚𝑖𝑛
50𝐸3 𝑚𝑖𝑛
50𝐺13 𝑚𝑖𝑛
50𝐺23
𝐷𝑛𝑛 = , 𝐷𝑠𝑠 = , 𝐷𝑡𝑡 =
𝑡𝑝 𝑡𝑝 𝑡𝑝
(5.3)

where 𝐸3 , 𝐺13 and 𝐺23 are the elastic modulus of the lamina and 𝑡𝑝 is the

thickness of a single ply. The values of the material parameters used are given

in Table 5.1.

Delamination onset is modeled using the quadratic failure criterion:

2
𝜎𝑛 2 𝜎𝑠 2 𝜎𝑡
( 𝑐) + ( 𝑐) + ( 𝑐) = 1
𝜎𝑛 𝜎𝑠 𝜎𝑡 (5.4)
Chapter 5 113

in which 𝜎𝑛 , 𝜎𝑠 and 𝜎𝑡 are the normal and shear stresses on the interface, and

𝜎𝑛𝑐 , 𝜎𝑠𝑐 and 𝜎𝑡𝑐 are the normal and shear strengths of the interface [184]. Once

Eq. (5.4) is satisfied, delamination is initiated. Then, interface damage evolution

is specified based on a fracture energy criterion and the bilinear softening law:


∫ 𝜎eff 𝑑𝛿eff = 𝐺𝐶𝑚
0 (5.5)

in which 𝜎eff and 𝛿 are the effective traction and displacement, respectively

[185], and 𝐺𝐶𝑚 is the mixed-mode fracture energy which is determined by the

Benzeggagh and Kenane (B-K) [186] relation as

𝐺𝐼𝐼 𝜂
𝐺𝐶𝑚 = 𝐺𝐼𝐶 + (𝐺𝐼𝐼𝐶 − 𝐺𝐼𝐶 ) ( )
𝐺𝑇 (5.6)

where 𝜂 is a material property, 𝐺𝐼𝐶 and 𝐺𝐼𝐼𝐶 are the Mode I and II critical

fracture toughness, and 𝐺𝑇 = 𝐺𝐼 + 𝐺𝐼𝐼 . The material properties used in the

delamination failure criterion are given in Table 5.1.

5.2.2.3. Modeling of stitching yarns by beam elements

Basic assumptions:

According to the experimental observations, following assumptions are made

as a basis for the stitch model [187]:

A. The stitch yarns are cylindrical with circular cross section.

B. The stitch yarns only carry tensile load.

C. The failure of stitch yarns is brittle like failure (fiber failure) under

tension.
114 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

D. The breakage of stitch yarns is assumed to occur at the delamination

plane.

E. No slippage occurs at the embedded stitches (i.e. no stitch pull out).

F. The tensile properties of the stitch yarns are obtained from the tensile

test of impregnated stitch yarns.

The SEM of a delaminated surface of the flax yarn stitched UD and woven flax

fiber composite associated with the assumptions (D) and (E) are shown in

Figure 5.5.

(a)

Fractured flax
stitch yarn
UD flax fibers
no slippage

(b)

Fractured flax
stitch yarn no slippage

Figure 5.5. SEM of the delaminated surface for the DCB test specimen of (a) UD and (b)
woven flax fiber composite. Fracture of stitch (flax) yarn occurs at the delamination surface
with minimum fiber pull-out.
Chapter 5 115

The crack bridging traction of through-the-thickness stitch reinforcements is

modeled using one-dimensional (1D) 2-node beam elements with a brittle

fracture response implemented in a UMAT subroutine. A maximum

longitudinal stress failure criterion was used to determine fiber failure under

tension. The material properties used in the modeling of the DCB test are

summarized in Table 5.1.

In this study, the material properties of the flax fiber laminate and its

interlaminar fracture toughness were empirically determined from the tensile

and DCB test results and the elastic properties of the glass fiber laminates were

adapted from literature [151].

Table 5.1. Material properties used in the numerical model of DCB test.

Glass fiber composite [151]

E11 = 47.79 GPa; E22 , E33 = 13.6 GPa; G12 , G13 = 5.89 GPa; G23 = 5.23 GPa

𝑣12 , 𝑣13 = 0.27; 𝑣23 = 0.3

UD flax fiber composite

E11 = 25.3 GPa; E22 , E33 = 4.1 GPa; G12 , G13 = 3.4† GPa; G23 = 2.8† GPa

𝑣12 , 𝑣13 = 0.25; 𝑣23 = 0.3†

𝐺𝐼𝐶 = 1.25 KJ/m2 ; 𝐺𝐼𝐼𝐶 = 1.57 KJ/m2 ; η=1

σC𝑛 = 35 Mpa; σ𝑆C = 50 Mpa

Stitch (flax) yarn

E11 = 38.97 GPa; 𝑋𝑠𝑡𝑖𝑡𝑐ℎ


𝑡 = 330MPa; 𝑅 𝑠𝑡𝑖𝑡𝑐ℎ∗ = 0.32 mm

† Estimated values.
∗ Stitch cross − sectional radius.
116 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

5.3. Modeling of in-plane fiber bridging by means of


cohesive law
In this Section, the trilinear cohesive law of Figure 5.4, which is being obtained

from superposition of two linear softening law, is used to model the

delamination crack propagation in the presence of fiber bridging (mechanisms

I and II). The semi-analytical approach proposed by Airoldi et al. [151] is

applied to the DCB test results (R-curves) of the unstitched UD flax fiber

composite to determine the parameters of the superposed cohesive laws.

5.3.1. Testing for fracture properties of unstitched composite

The GIC values obtained from the DCB tests of the unstitched UD flax fiber

composites in relation to the change in crack length, ∆a, during crack

propagation, are presented in Figure 5.6. In this figure, in addition to the

Modified Beam Theory (MBT) method, the GIC values determined from the

Compliance Calibration (CC) and the Modified Compliance Calibration (MCC)

methods [173] are also plotted to check for differences in the calculation of GIC

values through the use of different data reduction methods. The procedures of

applying these methods to the test data are described in Appendix A. It was

observed that these three methods, which are all based on the relation of Irwin-

Kies for strain energy release rate [188], led to a very similar GIC values with a

maximum variation of 2.5% for all stitched and unstitched natural fiber

composites (also see Table B.3 in Appendix B). Since the analysis is robust and

the MBT analytical corrections, as the most conservative method, work just as

well for the stitched composites, the GIC values reported and used for simulation

in this thesis are all calculated based on the MBT method.


Chapter 5 117

From the DCB test results of the unstitched flax composite, it was observed that

the interlaminar fracture toughness rose steadily with increasing (delamination)

crack length as the crack propagated in steady-state. In general, such a response

which is often referred to as the resistance curve or R-curve, indicates the

involvement of different physical phenomena occurring in tandem during the

fracture process. The R-curve association with fiber bridging is a well-

established phenomenon [189] and the slow rise in fracture resistance of the

flax/epoxy specimens up to the first peak G IC value at ∆a  20-30 mm

(Figure 5.6) agrees with our experimental finding of a large natural fiber

bridging zone that visually measured up to 25-30 mm in length when it was

fully developed (Figure 5.7). The R-curve effect is attributed to the increase in

apparent stiffness and delamination resistance of the specimen as a result of

fiber bridging.

This large scale fiber bridging is believed to be a unique characteristic of

untwisted UD natural fiber composites, as the preform architecture allows the

elementary natural fibers to behave in a way similar to discontinuous aligned

long fibers. The long bridging zone could also explain why the interlaminar

fracture toughness of the flax composite is significantly higher than those

measured for glass fiber composites (671 J/m2) [190] (see Chapter 4.6). It is

therefore essential to consider the effect of fiber bridging in the FE delamination

model for natural fiber composites, in order to achieve a more accurate

prediction of their strength and fracture behavior.


118 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

1600

1400 Propagation GC

1200

1000
GIC (J/m2)

800 Initiation G1

600

400
Visual onset Propagation (MBT)
200 Deviation from linearity Propagation (CC)
5% offset Propagation (MCC)
0
0 10 20 30 40 50 60

Change in crack length, ∆a (mm)

Figure 5.6. Experimental G IC values of an unstitched UD flax/epoxy composite calculated


using the MBT, CC and MCC methods proposed under ASTM D5528 [31].

The experimental results show that the average initiation value of GIC for the

UD flax/epoxy composite is 0.771 kJ/m2 and the steady-state GIC value is 1.25

kJ/m2 which is reached after approximately 25-30 mm of crack propagation.

The length of crack propagation before the steady-state phase of Mode I critical

strain energy release rate is estimated based on the mean value of the steady-

state GIC with a 95% confidence interval (for upper and lower limit).

It has been established for synthetic fiber reinforced composites that Mode I

delamination, which is assumed to propagate in the homogeneous resin-rich

region in-between fiber layers, can be modeled by a bilinear traction-separation

law (cohesive zone model) [191]. However, this model is not able to represent

the GIC R-curve when delamination is accompanied by extensive fiber bridging.

To address this, several authors have shown that the traction-separation law with
Chapter 5 119

non-linear softening can be used to model the R-curve response as well as the

enlarged process zone associated with crack bridging [150,152,153,155].

Fiber bridging zone

Figure 5.7. In-plane fiber bridging across the delamination plane of UD flax fiber composite
during DCB test.

In the next section, the trilinear cohesive law, which is being obtained from the

superposition of two linear softening law, is used to model the delamination

crack propagation in the presence of fiber bridging.

5.3.2. Superposition of cohesive laws for modeling the effect of

fiber bridging

A trilinear cohesive law (bilinear softening) of Figure 5.4, obtained from

superposing of two bilinear cohesive laws, is applied to model the fiber bridging

effect encountered during Mode I fracture of UD flax fiber composites.

Accordingly, a semi-analytical equation for extracting the superposition

parameters of the cohesive laws from the experimentally obtained R-curves are

used [151]. In this method, two bilinear cohesive laws with their own specific
120 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

characteristics are superposed to obtain a trilinear cohesive law. The

superposition parameters of m and n (Figure 5.4) are defined as follows:

𝐺1 = 𝑚𝐺𝑐 , 𝐺2 = (1 − 𝑚)𝐺𝑐
𝜎1 = 𝑛𝜎𝑐 , 𝜎2 = (1 − 𝑛)𝜎𝑐
(5.7)

where 𝜎 and 𝐺 refers to the cohesive strength and Mode I fracture toughness,

respectively. According to the experimental R-curve of the UD flax fiber

composites (Figure 5.6), the average initiation value of Mode I critical strain

energy release rate (𝐺1 ) was 0.771 kJ/m2 and the steady-state value (𝐺𝑐 ) was

1.25 kJ/m2 which was reached after approximately 25-30 mm of crack

propagation. Therefore, using the first relation of Eq.(5.7), the parameter m can

be calculated as follows

m = 𝐺1 ⁄𝐺𝑐 = 0.616
(5.8)

5.3.2.1. Calculation of the superposition parameter n

In order to determine the value of the parameter n, it is necessary to establish a

relationship between the characteristic length of the process zone for the

trilinear cohesive law (𝑙𝑐_𝑆𝑢𝑝 ) and the characteristic length of each primary

bilinear cohesive laws, 𝑙𝑐𝑖 , 𝑖 = 1, 2.

There are special cases in the relation between the characteristic length of the

process zone for the trilinear and the primary bilinear laws. These special case

are as bellow:

I. The sum of two bilinear laws is a bilinear law when 𝑚 = 𝑛.


Chapter 5 121

II. If one of the two bilinear cohesive laws has no associated fracture

toughness (i.e. either 𝑚 = 0 or 𝑚 = 1 ), the contribution of that

particular cohesive law is neglected.

III. If the strength of a bilinear cohesive with a nonzero fracture toughness

tends to zero (i.e. either 𝑛 → 0 or 𝑛 → 1), the superposition process

zone will tend to infinity.

IV. The order of the superposition of the two bilinear laws is irrelevant.

These special cases for the length of process zone of the trilinear cohesive law

(𝑙𝑐_𝑆𝑢𝑝 ) can be stated as follows

𝑙𝑐_𝑆𝑢𝑝 (𝑛, 𝑛) = 𝑙𝑐
𝑙𝑐_𝑆𝑢𝑝 (𝑛, 𝑚 = 0) = 𝑙𝑐2
𝑙𝑐_𝑆𝑢𝑝 (𝑛, 𝑚 = 1) = 𝑙𝑐1
lim 𝑙𝑐_𝑆𝑢𝑝 = ∞
𝑛→0 𝑜𝑟 1

𝑙𝑐_𝑆𝑢𝑝 (𝑛, 𝑚) = 𝑙𝑐_𝑆𝑢𝑝 (1 − 𝑛, 1 − 𝑚) (5.9)

The characteristic length of the process zone for the trilinear cohesive law

(𝑙𝑐_𝑆𝑢𝑝 ), based on the superposition parameters (𝑚, 𝑛), can be calculated using

the following interaction equation [151]

𝑚2 1 − 𝑚 2 𝑚(1 − 𝑚)
𝑙𝑐_𝑆𝑢𝑝 (𝑛, 𝑚) = [ 2 + ( ) − ]𝑙
𝑛 1−𝑛 𝑛(1 − 𝑛) 𝑐 (5.10)

in which all the conditions of Eq.(5.9) are satisfied. In Eq.(5.10), 𝑙𝑐 is the

material characteristic length which is an intrinsic fracture property of material.

The characteristic length 𝑙𝑐 is usually estimated as [151]


122 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

𝐸′𝐺
𝑙𝑐 = 𝛾 (5.11)
𝜎𝑐2

where 𝐺 and 𝜎𝑐 is the fracture toughness and strength of the material,

respectively. The modulus 𝐸 ′ under plane stress condition is defined as

−1

𝐸22 𝐸22
𝐸 ′ = 2𝐸22 (√2 [√ − 𝑣21 ] + )
𝐸11 𝐺12
(5.12)

where 𝑣21 is the in-plane Poisson’s ratio, 𝐸22 and 𝐺12 are the transvers Young

modulus and shear modulus, respectively. In Eq. (5.11), γ is a non-dimensional

parameter and depends on the damage/yielding process.

Generally, the length of process zone under steady-state propagation is


𝑠𝑠
estimated by the characteristic length (𝑙𝑝𝑧 ≈ 𝑙𝑐 ). However, the estimation is

valid only when the material characteristic length 𝑙𝑐 is smaller than the

structural dimension. This is because the value of γ in Eq. (5.11) is noticeably

influenced by the structural dimension. Airoldi et al. [151] proposed the

empirical relation of Eq.(5.13) in order to consider the effect of structural


𝑠𝑠
thickness in prediction of the steady-state process zone length, 𝑙𝑝𝑧 ,

𝛽
𝑠𝑠
𝑡𝑎𝑟𝑚
𝑙𝑝𝑧 =( ) 𝑙𝑐
𝑡𝑎𝑟𝑚 + 𝑙𝑐
(5.13)

where 𝑡𝑎𝑟𝑚 is the block thickness (thickness of plies stacking with same

orientation), and β is a non-dimensional parameter. They also conducted a

numerical analysis to assess the value of γ and β, and showed that when 𝛾 =
Chapter 5 123

0.70 and 𝛽 = 0.71, the best prediction of the process zone length was achieved

𝑠𝑠
for various rage of process zone lengths (𝑙𝑝𝑧 ).

Thus, Eq. (5.13) can be used to applied thickness correction on the characteristic

length obtained from Eq. (5.10) for the superposed cohesive law, 𝑙𝑐_𝑆𝑢𝑝 , as

𝛽
𝑠𝑠
𝑡𝑎𝑟𝑚
𝑙𝑝𝑧_𝑆𝑢𝑝 =( ) 𝑙𝑐_𝑆𝑢𝑝
𝑡𝑎𝑟𝑚 + 𝑙𝑐_𝑆𝑢𝑝
(5.14)

Nonlinear Eq. (5.14) can be solved for 𝑙𝑐_𝑆𝑢𝑝 , in which 𝑡𝑎𝑟𝑚 = 2.99 𝑚𝑚 and the

𝑠𝑠
steady-state process zone lengths is 𝑙𝑝𝑧_𝑆𝑢𝑝 = 30 𝑚𝑚 (estimated form

experimental R-curve). Finally, using the calculated values of parameter m and

𝑙𝑐_𝑆𝑢𝑝 , the nonlinear Eq. (5.10) can be solved for the parameter n iteratively,

which results in

𝑛 = 0.9699
(5.15)

The results of the FE analysis of DCB specimen using the superposition

parameters 𝑛 =0.9699 and 𝑚 = 0.6160 are presented in the next section.

5.3.3. FEM modeling of DCB test of unstitched composite

The trilinear cohesive law using the calculated values of 𝑛 and 𝑚 were applied

to the cohesive elements of the FE model of the DCB specimen. The force-

displacement response of the DCB test specimen model of unstitched UD flax

fiber composite using bilinear and trilinear cohesive laws, compared to the

experimental data, are shown in Figure 5.8.


124 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

It can be clearly seen that the prediction with the linear softening law is not

suitable for modeling the interlaminar fracture of the UD natural fiber composite.

When the initiation value of GIC is used (denoted by ‘FEM-Bilinear law (GIC

=0.77)’), the delamination onset is captured with good accuracy. However, the

model is not able to predict the force response during delamination propagation

and the overall force response is underestimated. On the other hand, Using the

steady-state GIC value (denoted by ‘FEM-Bilinear law (GIC =1.25)’)

overestimates the failure loads in the early period of crack growth that most

likely coincides with the development of the fiber bridging zone. Nevertheless,

it is able to capture the delamination propagation response in the presence of

fiber bridging traction which acts over the extended damage zone in the wake

of the crack tip. These results thereby confirm that the use of a bilinear law to

model delamination in the presence of extensive fiber bridging (Figure 5.7), can

introduce severe inaccuracy to the prediction of failure loads during DCB test.

Conversely, the use of the trilinear cohesive law (denoted by ‘FEM-Trilinear

law (GIC =1.25)’) gave a good approximation to the experimental load-

displacement curve. The nonlinear softening law is able to predict the onset of

nonlinearity of the loading curve corresponding to the physical onset as well as

the propagation of delamination. Generally, in the presence of fiber bridging,

the damage process zone is extended by inducing bridging traction over the

wake of the crack tip which leads to significant increase in the crack growth

resistance of fiber reinforced composites. Therefore, the use of nonlinear

softening law is necessary to differentiate decohesion of the matrix at the crack

tip and bridging traction in the crack wake.


Chapter 5 125

90

80

70

60

50
Load (N)

40

30

20
Exp #1 FEM-Bilinear law (GIC=0.77)
10 Exp #2 FEM-Bilinear law (GIC=1.25)
Exp #3 FEM-Trilinear law (GIC=1.25)
0
0 5 10 15 20 25 30 35 40 45
Cross-head displacement (mm)

Figure 5.8. Force–displacement response of DCB test specimen of unstitched UD flax fiber
composite; FEM & Experimental results.

The experimental R-curve is compared with the predicted results from the FE

analysis in Figure 5.9. Similar to the experimental data reduction method (i.e.

MBT), the Irwin-Kies equation (see Chapter3.9) was used to calculate the Mode

I critical strain energy release rate, GIC, of the FEM analysis.

The numerical results indicate that the trilinear law is able to predict the

initiation and the final steady-state values of experimental GIC very well.

However, the experimental R-curve hits a peak GIC before reaching steady state

propagation which is not predicted in the FE analysis results. This is because

the trilinear cohesive law parameters of 𝑚 and 𝑛 were calculated based on the

average steady-state GIC value of 1.25 kJ/m2. As discussed previously, this

value is chosen so as to cover a representative set of steady-state data points


126 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

from the experimental R-curve. It was shown by Sorensen [150] that the shape

of R-curve, i.e. the relation between the GIC and the crack length, was not

material property and depended on specimen geometry once large-scale fiber

bridging occurs across the delamination plane. However, the initiation and

steady-state value of Mode I critical strain energy release rate (GIC) were found

to be independent of specimen geometry.

The trilinear cohesive law obtained by this method was therefore employed as

the interlaminar response of the parent laminate to model the delamination of

stitched laminates.

1600

1400

1200

1000
GIc (J/m2)

800

600

400 Experimental
FEM-Bilinear law (G=0.77)
200 FEM-Bilinear law (G=1.25)
FEM-Trilinear law (G=1.25)
0
0 10 20 30 40 50 60
∆a (mm)

Figure 5.9. Experimental R-curve and FEM analysis results of bilinear and trilinear cohesive
law.
Chapter 5 127

5.4. Results of FEM modeling of DCB testing of stitched


composite
The force-displacement response of the numerical simulation of the stitched UD

flax fiber composite is plotted against the experimental results in Figure 5.10.

As discussed in Chapter 4, the DCB tests of the stitched UD composite exhibit

unstable crack growth followed by arrest, giving rise to stick-slip response in

the load-displacement curves. Here, the experimental curves are intended to

provide an indication of the dispersion of the onset of the stick-slip crack

propagation due to variation in the properties of stitch yarns (see Table 4.1) as

well as the stitching distribution.

As it can be seen, the FEM model shows reasonably good agreement with the

experimental responses. Although the first stick-slip onset fracture point is

different, the predicted response in terms of the peak and crack arresting loads

as well as the linear behavior of the load-displacement curve, are within the

range of the experimental results. The slope is characteristic of DCB specimens

due to the change in arm compliance as the effective arm length increases. The

regularity of the stick-slip crack growth in both numerical and experimental

results do suggest that the behavior is intrinsically linked to the strain energy

build up and release events induced by the stitch reinforcements.


128 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

140
Exp #1

120 Exp #2
Exp #3
Exp #4
100
FEM

80
Load (N)

60

40

20

0
0 10 20 30 40 50 60
Cross-head displacement (mm)

Figure 5.10. Force–displacement response of DCB test specimen of stitched UD flax fiber
composite; Experimental versus FEM analysis results.

A number of explanations can be offered for the discrepancy between the

experimental and predicted force-displacement response. First, some variation

in thickness of the DCB specimen arms which can occur with the resin infusion

manufacturing method leads to some variations in the initial stiffness of the

specimens (before crack initiation).

Second, the discrepancy in the stick-slip responses is also attributable to the

small local variations in the stitch punching locations and distribution. In

practice, actual stitch parameters (stitch length and row spacing) do deviate

from the nominal values used in the numerical simulation (see Figure 5.1) due

to hand-stitching of the fiber preform.


Chapter 5 129

Third, the variation could also have arisen from the properties assigned to the

stitch elements, as the consistency and properties of natural fibers can often vary

by up to 25-30%. A parametric sensitivity study of tensile properties of the stitch

fibers is given in next section. However, despite these discrepancies, the

predicted response is considered to be reasonably representative.

Figure 5.11 compares the predicted Mode I critical strain energy release rate of

the stitched DCB model with the experimental values of G IC with an identical

stitch areal fraction. It is noteworthy that the stitch areal fraction, despite the

plausible variation in the stitch distribution caused by hand-stitching, remains

unchanged in all the test specimens.

The results of Figure 5.11 indicate that the prediction of the stick-slip initiation

GIC values by the FEM analysis do give a reasonable representation of the

average of the experimental values for the stitched composites. The virtual DCB

test also exhibits the R-curve response, ramping from initiation GIC =

1.3 (kJ⁄m2 ) to the constant stick-slip propagation values of GIC =

2.26 (kJ⁄m2 ), which agree with the average of the experimental results very

well. The average experimental values are GIC = 1.34 (kJ⁄m2 ) and GIC =

2.22 (kJ⁄m2 ) for initiation and stick-slip propagation respectively. It is now

confirmed experimentally and with FEM prediction that stitching with twist-

less flax yarn can enhance the average interlaminar fracture toughness from

GIC = 1.25 (kJ⁄m2 ) for unstitched laminates to GIC = 2.26 (kJ⁄m2 ) for the

flax yarn stitched composites – i.e. an improvement of 80%. Detailed DCB test

results for unstitched and stitched flax fiber composites are presented in Table

B.3 of Appendix B.
130 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

3000

2500

2000

1500
GIc (J/m2)

Exp #1

1000 Exp #2
Exp #3
Exp #4
500
FEM

0
0 5 10 15 20 25 30 35 40 45 50
Change in crack length, ∆a (mm)

Figure 5.11. Experimental and FEM analysis R-curve of stitched UD flax fiber composite.

The evolution of the crack length against the cross-head opening displacement

obtained by experimental and FEM simulation of un/stitched DCB test

specimens are compared in Figure 5.12. The model appears to be able to give a

satisfactory prediction of interlaminar failure of the unstitched and stitched

specimens. The fit between the predicted and experiment crack lengths for the

unstitched composite laminates is very good. This means that for any given

values of crack length, the numerical compliance (P/δ) accurately matches the

experimental values.

For the stitched model, despite the discrepancy between the crack arrest lengths,

the numerical model shows that the crack growth takes place at a similar rate to

the experimental data. As discussed above, the difference between the predicted

crack length and the experimental data (against the cross-head opening

displacement) is attributable to the discrepancy in stitch punching locations


Chapter 5 131

between the actual specimens and the numerical model as well as intrinsic

variation in the tensile properties of the stitch fibers. It is worth mentioning, due

to stick-slip crack growth, only the crack arrest length and the associated

maximum opening displacement are plotted for the stitched composites.

60

50
cross-head opening
displacement
Change in crack length, ∆a (mm)

40
∆a

30

20

10 Exp (Stitched) FEM (Stitched)

Exp (Unstitched) FEM (Unstitched)

0
0 5 10 15 20 25 30 35 40 45 50
Cross-head opening displacement (mm)

Figure 5.12. Change in crack length against cross-head opening displacement for experimental
and FEM analysis

Figure 5.13 compares the delamination growth threshold markings on the

numerical and experimental load-displacement curves of the stitched

composites (i.e. point 1, 2 and 3). As can be seen, there is a good agreement

between the calculated and the measured curves. The numerical model is able

to accurately predict the bilinear rise in the force in the elastic region, the peak

force as well as the force jagged pattern due to the stick-slip crack growth.
132 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

(a) Experimental (b) Numerical


140 140
Unstitched (Exp) Stitched (FEM)
120
 Stitched (Exp) 120  Unstitched (FEM)

100 100

  

Load (N)
80 80
Load (N)


60 60

40 40

20 20

0 0
0 10 20 30 40 50 0 10 20 30 40 50
Cross-head displacement (mm) Cross-head displacement (mm)

Figure 5.13. The 3-step delamination growth marking on the load-displacement curve of DCB
test of stitched composite; comparison of prediction with experimental results.

In the DCB test, the opening force of the specimen’s arms is linear till it reaches

the point 1, as shown in Figure 5.13. This point indicates the onset of

delamination from the end of the PTFE insert (pre-crack) in both the unstitched

and stitched composites (Figure 5.14a). Beyond point 1, unlike the stable crack

propagation in the unstitched composite, the linear force rise continues in the

stitched composite while the gradient is slightly changed compared to the earlier

part. At this step, the crack passes round the stitch fiber and is arrested behind

the stitch line while the stitch fibers are bridging the crack, as shown in the

damage evolution of the cohesive layer of the DCB test model in Figure 5.14b.

The opening force then rises to a maximum value at point 2, where the failure

stress of stitch elements is reached. Finally, the stitch elements fracture and the

force drops abruptly till stops at point 3, while the force is always above the

failure force of the unstitched composite. Releasing the stored strain energy in

the stitch elements causes the crack to propagate at higher speeds towards the

vicinity of the next stitch line, as shown in Figure 5.14c. This progressive failure
Chapter 5 133

is repeated for each stitch line and leads to the unstable stick-slip crack growth

behavior observed experimentally.

(a)

(b)
134 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

(c)

Figure 5.14. Interlaminar damage (delamination) evolution within the stitched UD flax fiber
composite; (a) the cohesive interface failure propagates and passes round the stitch elements,
(b) the crack then is arrested behind the stitch elements while the stitch elements are stretched,
(c) the stitch elements fractures and the stored energy is released to propagate the cohesive
crack at higher speed until the next row of stitch elements.

5.4.1. Parametric study

It was discussed in the previous section that variation in the tensile properties

of stitch yarn, i.e. flax yarn, can explain the discrepancy between the predicted

and measured values of the Mode I interlaminar fracture toughness of the

stitched composite. It is rather clear that the strength of the stitch material is the

most important parameter and have a significant influence on the bridging

traction, and on the delamination resistance accordingly. In this regards, a

parametric study was performed to determine the effect of the tensile strength,

Xtstitch , of the stitch yarn on the delamination resistance of the stitched UD flax

fiber composite.

The experimental study on the tensile properties of the flax yarn showed a

coefficient of variation of 25-30% for the tensile strength of the impregnated

flax yarn (see Table 4.1). Here, we consider two extreme values for the strength

of the stitch elements in which they are 15% above and below the average value
Chapter 5 135

of the strength used in FEM analysis in previous section (i.e. Xtstitch =

380 MPaand Xtstitch = 280 MPa for the maximum and minimum values of the

strength, respectively).

The effect of variation of stitch strength on the load-displacement response and

the interlaminar fracture toughness are shown in Figure 5.15a and b,

respectively. As can be seen, the variation of the strength of stitch elements has

noticeable effect on the opening load as well as the interlaminar fracture

toughness, GIC. A variation of 30% in the strength of stitch elements leads to

20% and 31% change in the peak force and the propagation value of GIC,

respectively. This is an indication of the sensitivity of Mode I interlaminar

fracture toughness to the bridging traction taking place across the delamination

plane. In other words, any variation in the strength of the stitch material would

be reflected in the toughness adding to the composite. It is apparent from the

Figure 5.15b that the majority of the experimental data fall between the

numerically predicted upper and lower limits. As a result, it can be said that the

dispersion in the experimentally measured values of GIC is mainly attributable

to the variation in the strength of the stitch yarn.


136 Numerical modeling of Mode I interlaminar failure of stitched flax fiber composites

(a)
140

120

100
Force (N)

80

60

40
Main
Stitch Strength = 330 Mpa (average)
20 Stitch
25% X Strength = 380 Mpa (+15%)

StitchXStrength = 280 Mpa (- 15%)


_25%
0
0 5 10 15 20 25 30 35 40 45
Cross-head displacement (mm)

(b)
3000

2500

2000
GIc (J/m2)

1500

1000 Experimental
Experimental
Stitch Strength = 330 Mpa (average)
Main

500 Stitch
25% X Strength = 380 Mpa (+15%)
_25%
StitchXStrength = 280 Mpa (- 15%)

0
0 5 10 15 20 25 30 35 40 45 50
Change in crack length, ∆a (mm)

Figure 5.15. Effect of variation of stitch strength; (a) Force–displacement response and (b) R-
curve of the stitched UD flax fiber composite.

5.5. Conclusion
The effect of introducing out-of-plane flax fiber stitches on the Mode I

interlaminar fracture toughness of the UD flax/epoxy composite laminates was

experimentally and numerically investigated. The DCB test results for the

unstitched flax fiber composites showed a strong R-curve effect which is


Chapter 5 137

attributed to the large-scale in-plane fiber bridging, observed during

delamination advancement. The experimental findings also showed that the use

of twist less flax yarn stitches was effective and can improve the Mode I critical

strain energy release rate at by around 78% to over 2 kJ/m2.

In the numerical analysis, a cohesive element with trilinear cohesive law

obtained from the superposition of two bilinear cohesive laws was used to

model the R-curve effect. It was shown that this approach is accurate enough to

capture the initiation and steady-state propagation values of GIC of the R-curve

induced by the large scale in-plane fiber bridging. The resultant cohesive

interface was then superposed with beam elements as the stitch fibers to

simulate the delamination of stitched laminate in the presence of extensive in-

plane fiber bridging. The stitch model was built based on experimental

observations in which the fracture of stitches was found to occur at the

delamination plane with no slippage effect at the embedded end. The FE

analysis results showed reasonably good agreement with the experimental

results and the prediction of the Mode I critical strain energy release rate for the

initiation and propagation phase of the delamination were within the range of

the measured values. Moreover, the parametric study on the strength of the

stitch yarn showed that the variation in the measured values of G IC was in line

with the dispersion of strength of the stitch yarn. The results of this chapter

suggest that stitching is a promising technique to enhance the interlaminar

fracture toughness of flax/epoxy composite laminates, and it is crucial to

consider the effects of in-plane fiber bridging to have an accurate

characterization of GIC values for stitched natural fiber composites.


138 Low-velocity impact performance of stitched flax/epoxy composite laminates

Chapter 6
Low-velocity impact performance of

stitched flax/epoxy composite laminates

6.1. Introduction
Low-velocity transverse impact performance of composites relates to how well

the material can stand up to damages caused by tool drops and low impact strike

events that generally occur in an operating environment. Its relevance and

importance for traditional carbon/glass composites arises from observations that

low energy impact damages typically manifest in the form of sub-surface

delamination which makes surface detection difficult [13,14]. Although natural

fiber composites may not withstand the same level of impact forces,

understanding its low-velocity transverse impact behavior and damage

development remains a key requirement for its design, damage detection, and

maintenance.

The present chapter is devoted to the understanding of the low-velocity impact

performance of the optimally-stitched flax fiber composites developed in

chapter 4. Drop-weight impact test was performed to investigate the behavior

of the unstitched and stitched woven flax/epoxy composites as well as cross-ply

flax/epoxy composites under low-velocity impact loads in accordance to ASTM


Chapter 6 139

standard D7136 (3.10). Low-velocity drop weight impact tests are conducted at

two different energy levels, 19.6J and 44.6J, to attain both non-perforation and

full-perforation impact conditions, for the study of impact damage resistance

and impact energy absorption behavior of the composites empirically.

6.2. Specimen types and the drop-weight impact test


parameters
It was presented earlier that an optimum areal fraction of stitch should be

considered when using natural fiber stitches to strike a balance between the

improvement in interlaminar fracture toughness against the reduction in tensile

properties. Again, twist-less flax yarn and twisted cotton thread were used to

stitch together four layers of the flax woven fabrics at an equivalent and

optimum stitch areal fraction of close to 0.01. Unstitched cross-plies [0/90]4s of

continuous flax fibers are also manufactured to a similar thickness for

benchmarking. The plain weave specimens are denoted PW0 for unstitched,

PWC for cotton-thread stitched and PWF for flax yarn stitched, respectively.

UD0 refers to the unstitched cross-ply specimens.

Stitch parameters applied to the woven preforms in the warp (X) direction were:

𝑆𝑅 = 4𝑚𝑚 (stitch row spacing) and 𝑆𝐿 = 3𝑚𝑚 (stitch length) for PWC; and

𝑆𝑅 = 𝑆𝐿 = 8𝑚𝑚 for PWF. These parameters were chosen in order to achieve

an equivalent and near optimum value of the stitch areal fraction as defined by

Eq.(3.1), for flax yarn and cotton thread based on their cross-section area given

in Table 4.1. Details concerning the composites’ fiber preform are presented in
140 Low-velocity impact performance of stitched flax/epoxy composite laminates

Table 6.1. The laminate thicknesses were highly consistent, measuring 4±0.10

mm with a coefficient of variation of less than 2% for each specimen. Test plates

of 100 mm by 150 mm were cut via water-jet and oven dried immediately at 60℃

for at least 24 hours before the impact tests.

Table 6.1. The composite lay-up, notation and stitch parameters used for the manufactured
flax fiber laminates for the drop weight impact test.

Areal density Stitch Areal


Flax Lamina Laminate Stitch 𝑽𝒇
of lamina Layup Fraction
preform nomenclature material (%)
[g/m2] x10-2
PW0 [0]4 - 0
4×4Plain
PWC 500 [0]4 Cotton 0.92 31±2
Weave
PWF [0]4 Flax 1.03
Unidirectional UD0 110 [0/90]4s - 0 40±2

The impact force-time history was obtained from a force sensor mounted on the

impactor, and the contact velocity was measured from the images recorded by

a high-speed camera filming at 15000 frame/second (see Section 3.10). After the

impact test, all the specimens were visually checked for any external damage

on both impacted and back surfaces. The damage type, as well as length of the

cracks, were recorded for all the specimens.

The following sections shall present the impact test study results, starting with

the post-impact visual inspection the test pieces in Section 6.3 to first understand

the different fracture locations and behavior of natural fiber composites,

followed by the discussion on the force-time and force-displacement results in

Sections 6.4 and 6.5 respectively. Lastly, the energy absorption and impact

fracture toughness will be covered in Section 6.6.


Chapter 6 141

Table 6.2. Information of Impact test and specimen’s thickness.

Impact Impact velocity Average thickness Specific impact


sample
Energy (J) (m/s) (mm) energy† (J/mm)

PW0 4.01 0.05 4.89 0.01


PWC 3.92 0.08 5.00 0.03
19.6 1.98
PWF 3.99 0.04 4.91 0.01
UD0 3.97 0.03 4.94 0.01
PW0 4.06 0.06 10.99 0.04
PWC 4.01 0.04 11.12 0.02
44.6 2.98
PWF 3.98 0.10 11.21 0.07
UD0 3.99 0.10 11.18 0.07
† Represents the ratio of impact energy to specimen thickness.

6.3. Visual inspection of post-impact specimens


Impact resistance is defined as the ability of a material to resist permanent

changes (i.e. damage) due to a loading event beyond the design load. In this

study, the impacted specimens are visually inspected to identify the external and

visible damages, so as to understand the types and positions of the cracks that

had developed.

6.3.1. Non-perforation impact test (19.6J)

Figure 6.1 shows both impact and bottom face of woven and cross-ply flax

specimens subjected to 19.6J impact energy. It was observed that for all woven

flax fiber laminates (PW0, PWC, and PWF), a cross-shaped crack formation

had developed along the weft and warp directions at both the impact and the

back surfaces of the specimens. The intersection of the longitudinal and

transverse cracks corresponded to the impact location. This suggests that natural

fiber breakage was predominant, unlike the stronger synthetic (carbon or glass)
142 Low-velocity impact performance of stitched flax/epoxy composite laminates

fibers which could generally resist fractures at this energy level and exhibit

mainly matrix failures [192,193].

Comparison between the unstitched and stitched specimens showed that the

fracture mechanisms were more or less identical. However, crack lengths in

both directions were longer for the stitched specimens. The overall crack length

was measured to be 27±2% (PWC) and 41±2% (PWF) longer for cotton and

flax stitched laminates, respectively. This could be attributed to the higher

incidence of defects such as in-plane fiber damage and resin-rich spots caused

by stitching which facilitates the creation and propagation of intralaminar cracks.

There are also some effects arising from the stitch fiber dimension, which will

be discussed in Section 6.6.

For the cross-ply laminate (UD0), a combination of matrix failure (transverse

crack) and fiber breakage (longitudinal crack) formed a cross-shaped damage at

the back surface. Matrix dominated cracks along the top-ply fiber orientation

with sub-surface delamination were observed on the impact face of the

specimens. The visible peanut-shape delamination zone around the locus of the

impact was only observed for the UD0. This type of delamination damage is

usually observed in impacted synthetic fiber composites with cross-ply lay-up,

attributing to the interlaminar shear stress distribution in 0°/90° interface

produced by transverse loading of the laminate [194]. It was presented earlier

in Chapter 4.6 that the interlaminar fracture toughness of the UD laminate is

generally much lower than the woven laminates. Therefore, it is conceivable

that the cross-ply laminates should be more vulnerable to delamination under

transverse loading due lower fracture resistance. The effects of the delamination
Chapter 6 143

will be discussed more under in Section 6.4 together with the impact force

histories.

(a) (b) (c) (d)

Delamination
Transvers crack
Longitudinal crack

Warp 90°
Matrix failure

Weft 0°

PW0 – Impact face PWC – Impact face PWF – Impact face UD0 – Impact face

Fiber failure

90°
Matrix failure

PW0 – Back face PWC – Back face PWF – Back face UD0 – Back face

Figure 6.1. Damage of 19.6J impact (non-perforation) on both faces of (a) unstitched woven –
PW0, (b) cotton stitched woven – PWC and (c) flax stitched woven – PWF, (d) unstitched
cross-ply – UD0.

6.3.2. Perforation impact test (44.6J)

The perforated impact specimens tested at 44.6J are shown in Figure 6.2. The

shape and size of the damage to the woven laminates were fairly similar, with

no distinct differences between the unstitched and stitched woven laminates

found. The damage appeared to have been formed by matrix and fiber fractures

that emanated from the locus of impact along the principal fiber directions with

very little delamination detected, followed by flexural bending failures of the

quadrants created. On the other hand, the impact face of the cross-ply laminate
144 Low-velocity impact performance of stitched flax/epoxy composite laminates

(UD0) shows visible delamination besides fiber and matrix failures. Quadrant

bending failures and some minor delamination adjacent to the crack lines and

in the sub-surface plies was also observed from the bottom face. This meant that

the impact energy was dissipated by both major crack formation and

delamination in cross-ply architectures.

(a) (b) (c) (d)

Warp 90º 90º 90° 90º

Weft 0°
0º 0º 0º
PW0 – Impact face PWC – Impact face PWF – Impact face UD0 – Impact face

Warp

Weft

PW0 – Back face PWC – Back face PWF – Back face UD0 – Back face

Figure 6.2. Damage of 44.6J impact (non-perforation) on both faces of (a) unstitched woven –
PW0, (b) cotton stitched woven – PWC and (c) flax stitched woven – PWF, (d) unstitched
cross-ply – UD0.

6.4. Impact force- time history


The force-time history of the 19.6J and 44.6J impact tests are shown in

Figure 6.3 and Figure 6.4, respectively. The force-time responses were

smoothed using fast Fourier transform (FFT) with cut off frequency of 0.01 Hz.
Chapter 6 145

This frequency was chosen to filter the force oscillation introduced by the

natural frequency of the impactor and the test apparatus.

(a) (b)
3500 3500
PW0-#1 UD0-#1
3000 3000
PW0-#2 UD0-#2
2500 PW0-#3 2500
UD0-#3

Force (N)
2000 2000
Force (N)

1500 1500

1000 1000

500 500

0 0
0 5 10 15 20 25 0 5 10 15 20 25
Time (ms) Time (ms)

(c) (d)
3500 3500
PWC-#1 PWF-#1
3000 3000 PWF-#2
PWC-#2
PWC-#3 PWF-#3
2500 2500
Force (N)
Force (N)

2000 2000

1500 1500

1000 1000

500 500

0 0
0 5 10 15 20 25 0 5 10 15 20 25
Time (ms) Time (ms)

Figure 6.3. Impact force-time histories of 19.6 J (non-perforation) for (a) unstitched woven –
PW0, (b) unstitched cross-ply – UD0, (c) cotton stitched woven – PWC and (d) flax stitched
woven – PWF.
146 Low-velocity impact performance of stitched flax/epoxy composite laminates

(a) (b)
3500 3500
PW0-#1
UD0-#1
3000 PW0-#2 3000
UD0-#2
PW0-#3
2500 2500 UD0-#3
Force (N)

Force (N)
2000 2000

1500 1500

1000 1000

500 500

0 0
0 5 10 15 20 0 5 10 15 20
Time (ms) Time (ms)

(c) (d)
3500 3500
PWC-#1 PWF-#1
3000 PWC-#2 3000 PWF-#2
PWC-#3 PWF-#3
2500 2500
Force (N)
Force (N)

2000 2000

1500 1500

1000 1000

500 500

0 0
0 5 10 15 20 0 5 10 15 20
Time (ms) Time (ms)

Figure 6.4. Impact force-time histories of 44.6 J (perforation) for (a) unstitched woven – PW0,
(b) unstitched cross-ply – UD0, (c) cotton stitched woven – PWC and (d) flax stitched woven
– PWF.

The force-time histories for all types of laminates were highly consistent, and

the coefficient of variation of the impact force measurements were less than 5%.

Similar consistency was previously observed for the damage resistance of

specimens in terms of crack lengths and damage pattern. Therefore, although

large variations generally exist in the properties of natural fibers, consistent

composite properties can be achieved with thorough impregnation of fiber

preforms and a high-quality fabrication process.

It can be seen that both the stitched laminates (PWC and PWF) yielded a lower

force values compared to unstitched laminates. Among all the laminates, the
Chapter 6 147

cross-ply laminate (UD0) owns the highest force values. Moreover, the contact

duration in the non-perforation test was longer for both the stitched laminates,

indicating a larger damage development in the laminates. This duration was the

least for the UD0.

Some differences were also observed between the impact response of the woven

and the cross-ply laminates, which may be influenced by their impact fracture

mechanisms. The typical force-time history of woven and UD cross-ply flax

fiber laminates under drop-weight impact are shown in Figure 6.5a and b,

respectively. In the elastic region, the impact force increases linearly until it

reaches a characteristic force P1 at which a change in stiffness occurs due to

damage initiation. The value of P1 is therefore regarded as an indicator of the

laminate’s ability to resist damage initiation [81]. The values of P1 for the

composites studied are given in Table 6.3. It was found that the impact damage

initiation force (P1) for the stitched laminates at 19.6J was lowered by about 10%

due to weakening of the primary woven structure by stitch defects, such as resin

pockets and in-plane fiber distortion, which elevates the stress concentration

around stitch loops. Increasing the impact energy to 44.6J led to a slight increase

in the damage initiation force due to strain rate dependency [195]. As the

unstitched primary structure completely failed at 44.6J with full perforation by

the impactor, the structural weakening contributed by stitch defects became

relatively insignificant, and therefore the measured values of P1 were almost

identical for both the unstitched and stitched laminates.


148 Low-velocity impact performance of stitched flax/epoxy composite laminates

(a) (b)
3000 3500
Woven lamintes Cross-ply laminates
2500
PMax 3000 PMax
P1
P1 2500
2000

Force (N)
Force (N)

2000
1500
1500

1000
1000

500
500

0 0
0 5 10 15 20 0 5 10 15 20

Time (ms) Time (ms)

Figure 6.5. Typical representative curve of the force-time responses for (a) woven flax
laminates, (b) cross-ply flax laminate.

Beyond P1, the behavior of the impact force can be attributed to the damage

development within the laminate. It is seen from Figure 6.5 that there is an

apparent difference between the force response of woven and cross-ply

laminates, where a drop in the force right after P1 followed by a short period of

force perturbation was observed in the latter. This can be related to the

propagation of delamination cracks observed in the cross-ply laminate

(Figure 6.1), whereas there is no evidence of delamination in the woven

laminates under impact loading, further verified in Figure 6.6. It is noteworthy

that the cross-ply laminate’s force-time response in Figure 6.5b resembles those

of the conventional composites [129] where delamination is a dominant failure

mode under transverse impact load. The general response of conventional

composites subjected to low-velocity impact was discussed in Chapter 2.4.

It was shown in Chapter 4.6 that the delamination resistance of the woven flax

composite is noticeably higher than the UD flax composites [180]. The high

interlaminar toughness and relatively low in-plane strength of the woven flax
Chapter 6 149

fiber lamina can be a reason for having in-plane fiber dominated fractures prior

to the interlaminar damage development (delamination), as illustrated

schematically in Figure 6.7. In the cross-ply laminate, delamination led to a

rapid decrease in the force by reducing the bending stiffness of the laminate.

The force perturbation after P1 implies the energy dissipation by propagation of

the in-plane damages.

A small reloading to a maximum force of Pmax also happens in both the woven

and cross-ply because of strain rate dependency as there exist a sufficient

potential energy stored in the impactor [129]. The maximum impact force

values (Pmax) of the cross-ply laminates are higher than the (unstitched and

stitched) woven laminates (Table 6.3), suggesting that the unstitched UD cross-

ply fiber orientation offers the highest load resistance against impact damage

propagation for laminates of similar thicknesses. This can be largely attributed

to the more compact packing of cross-ply UD fibers within the laminate, which

resulted in higher fiber volume fraction ( 𝑉𝑓 = 0.4 for UD0 laminate) and

therefore higher tensile stiffness and strength over the lower 𝑉𝑓 woven

laminates [8].

For the woven composites which are all of similar 𝑉𝑓 (= 0.3), the stitched

specimens (regardless of the stitching fiber) always led slightly lower values of

Pmax. It has been well documented that stitching could disrupt the in-plane fiber

alignment and increase the matrix-rich regions between the woven fiber tows

[131]. Although stitch defects are expected to be less detrimental in natural fiber

composites due to the lower fiber strengths and higher discrepancy in the fiber

properties [180], their higher incidence will still facilitate crack formation and
150 Low-velocity impact performance of stitched flax/epoxy composite laminates

damage propagation at lower values of force under transverse impact loading.

This can also explain why the principal axis cracks are longer in the stitched

laminates (Figure 6.1a-c).

Unstitched Woven (Impact Face)

❶❷ ❸❹ ❺

❶ ❷ ❸ ❹ ❺

Figure 6.6. Through-thickness inspection of the impacted woven laminates by the cross-
sectional microscopy.
Chapter 6 151

σC

σT

Figure 6.7. Schematic view of a woven flax composite under impact loading in which the
failure in the fiber layers due to tensile (𝜎𝑇 ) or compressive (𝜎𝐶 ) stress happens prior to a
failure in resin-rich inter-ply layers due to shear stress (τ).

The similarity in impact forces at the two energies is attributed to the low

strength of the flax fiber reinforcements, resulting in fiber fracture being the

dominant failure mode. As can be seen from Figure 6.1 and Figure 6.6, the main

fracture travels pretty much straight through the thickness of the panel,

dominated by matrix and fiber fractures. Once the impact force reaches P1, the

impact energy starts to be dissipated by the damage mechanisms within the

composite, preventing a substantial increase in the impact force. This is one of

the main differences in characteristics between the impact fractures of natural

fiber composites and conventional glass or carbon fiber composites where in the

latter, delamination is the dominant failure mode [129]. The unbroken fibers in

the conventional composites would provide an elastic strain response to higher

impact loads.

6.5. Impact force-displacement response


The impactor force against displacement curves of the both impact energies for

all the specimens are illustrated in Figure 6.8 and Figure 6.9. The consistency

observed in the force-time histories and impact damages are also clearly evident
152 Low-velocity impact performance of stitched flax/epoxy composite laminates

in the force-displacement responses. The consistency in these results indicates

that with the right understanding of its failure mechanisms, the behavior of flax

composites under impact may be predicted by numerical modeling to be used

in design and development pursuits. This was shown to be true, and the model

developed for flax laminates is presented Chapter 7.

(a) (b)
3500 3500
PW0-#1 UD0-#1
3000 PW0-#2 3000 UD0-#2
PW0-#3 UD0-#3
2500 2500
Force (N)

2000 Force (N) 2000

1500 1500

1000 1000

500 500

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Displacement (mm) Displacement (mm)

(c) (d)
3500 3500
PWC-#1 PWF-#1
PWC-#2 PWF-#2
3000 3000
PWC-#3 PWF-#3
2500 2500
Force (N)

Force (N)

2000 2000

1500 1500

1000 1000

500 500

0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Displacement (mm) Displacement (mm)

Figure 6.8. Impact force-displacement response of 19.6 J for (a) unstitched woven – PW0, (b)
unstitched cross-ply – UD0, (c) cotton stitched woven – PWC and (d) flax stitched woven –
PWF.
Chapter 6 153

(a) (b)
3500 3500
PW0-#1
UD0-#1
3000 PW0-#2 3000
PW0-#3 UD0-#2
UD0-#3
2500 2500
Force (N)

Force (N)
2000 2000

1500 1500

1000 1000

500 500

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Displacement (mm) Displacement (mm)

(c) (d)
3500 3500
PWC-#1 PWF-#1
3000 PWC-#2 3000 PWF-#2
PWC-#3 PWF-#3
2500 2500
Force (N)
Force (N)

2000 2000

1500 1500

1000 1000

500 500

0 0
0 5 10 15 20 25 30 0 5 10 15 20 25 30
Displacement (mm) Displacement (mm)

Figure 6.9. Impact force-displacement response of 44.6 J for (a) unstitched woven – PW0, (b)
unstitched cross-ply – UD0, (c) cotton stitched woven – PWC and (d) flax stitched woven –
PWF.

For cross-specimen analysis, representative data from the impact tests of the

four composites are plotted together in Figure 6.10a and b. As can be seen, the

force-displacement curves for non-perforation impact are generally divided into

three zones of A, B, and C, as shown in Figure 6.10a [10]. Zone A represents

the elastic response of the specimen until the force has built up to the damage

initiation level. In zone B, the damage is initiated and continues to develop as

the impactor presses further into the plate until it ran out of energy and stopped

(motion reversal point). The force perturbations in this zone are associated with

the formation of fiber and matrix damages within plies, i.e. damage propagation.

Generally, a longer zone B is an indication of more damage development within


154 Low-velocity impact performance of stitched flax/epoxy composite laminates

the laminate and consequently larger damage zone. Finally, zone C (in the case

of non-perforation impact) is characterized by the force unloading part of the

test where the impactor rebounds with the help of stored elastic energy from the

plate until it loses contact with the specimen. The residual displacement at the

end of zone C indicates the permanent depression of the plate. For high energy

impact (perforation test) in Figure 6.10b, zone C is replaced by a rapid force

drop without strain recovery due to the full penetration of the impactor.

In the non-perforation impact tests shown in Figure 6.10a, the maximum

deflection (at the end of zone B) is almost 25-30% higher for the stitched

laminates than the unstitched PW0 and UD0 composites. This is despite the

observations of the bending stiffness at the elastic region (zone A) being almost

identical for all specimens. This could be associated with the larger damage size

(longer cracks) generated in the stitched laminates.

Increasing the impact energy to 44.6J led to laminate perforation and the

absence of impactor rebound in zone C. In this case, stitching appears to have

insignificant effect, and perforation occurred at nearly the same deflection for

the both unstitched and stitched laminates, in Figure 6.10b.


Chapter 6 155

(a)
3500
PW0
PWC

B
3000 PWF
UD0
2500

2000
Force (N)

1500 A
1000

C
500

0
0 2 4 6 8 10 12 14
Displacement (mm)

(b)
3500
PW0

3000 B PWC
PWF
UD0
2500
Force (N)

2000

1500
A
1000

500

0
0 5 10 15 20 25
Displacement (mm)

Figure 6.10. Different zones of impact force-displacement response for all the flax laminates
subjected to (a) 19.6J and (b) 44.6J impact energy.

6.6. Energy absorption and impact fracture toughness


The energy absorbed by the specimen with time was determined by calculating

the area under the force-displacement curve. This value of absorbed energy

refers to the amount of energy transferred to the specimen consisting of the

recoverable strain energy before the rebound, and the non-recoverable energy
156 Low-velocity impact performance of stitched flax/epoxy composite laminates

losses dissipated by the specimen through damage formation, friction, heat, and

sound. The energy absorption curves versus time for both 19.6J and 44.6J

impact energies tests are given in Figure 6.11.

50

Perforation impact 44.6 J


45

40

35

30
Energy (J)

25
Non-perforation impact 19.6 J
20

15 EAbsorbed

10

5
PW0 PWC PWF UD0
0
0 5 10 15 20 25 30

Time (ms)

Figure 6.11. Energy absorption curves under 19.6J and 44.6J impact energies.

At the non-perforation impact level (19.6J), the turning point of the energy

absorption curves at the maxima represents the moment that the impactor had

stopped and all energy of the impactor has been completely transferred to the

specimen. Thereafter, the elastic strain energy stored in the specimen is returned

to the impactor, causing it to rebound till separation and the energy curve at this

stage is characterized by a decreasing rate of decline. The final energy value

indicates the amount of energy dissipated (absorbed) by specimen largely

through damage development within the composite. It is worth mentioning that


Chapter 6 157

the energy of free vibration of the specimen is very small and is assumed to be

negligible due to the quasi-static condition of the low-velocity impact test

[129,196]. However, at the perforation impact level (44.6J), the energy of the

impactor is high enough to break through the specimen. Therefore, there is no

decline observed in energy absorption curve and most of the absorbed energy is

dissipated by the specimen through damage crack propagation and some

vibrational kinetic energy in the test piece. Moreover, as the impactor is still

traveling with unconverted potential energy within it at the moment of full

impactor penetration, the maximum energy received by the specimens will

always be less than the total energy of the impactor (i.e. 44.6J). This maximum

energy represents the impact perforation threshold energy of the laminates.

Having a lower perforation threshold means that the composite laminate

displays less resistance against the penetration of an impactor [87].

The ratio of the absorbed energy per total impact energy (Eabsorbed/Eimpact ,

expressed in percentage) for both non-perforation and perforation impact tests

are reported in Table 6.3. The variability in this ratio for the different types of

composite systems is due to the effects of stitch parameters, in-plane fiber

preform architectures, the fiber volume fraction as well as the variation in

laminate thickness. These inevitably influence the damage development and

energy absorption under impact conditions. At the non-perforation impact level

(19.6J), the ratio of Eabsorbed/Eimpact for the stitched woven composites were about

12-18% higher than that of the unstitched woven laminates, corresponding to

the longer crack lengths observed in the stitched woven laminates. This is

mainly because of the defects caused by stitching which lead to easy formation
158 Low-velocity impact performance of stitched flax/epoxy composite laminates

and development of damages in the stitched composites, as discussed in

Section 6.4. In the case of full perforation impact, the flax yarn stitched

laminates (PWF) and the cross-ply laminates (UD0) exhibited the lowest and

highest perforation thresholds, respectively. The higher impact resistance of the

cross-ply laminates (UD0) compared to the woven laminates can be attributed

to their higher fiber volume fraction ( 𝑉𝑓 = 0.4 ) [93,99], whereas the low

perforation resistance of the flax yarn stitched laminates is associated with the

larger defects induced by thicker yarn (flax) stitching.

Unlike conventional glass or carbon fiber reinforced composites where low

impact energy cracking tends to propagate through the matrix-rich inter-ply

planes between the fiber layers, the strengths of the woven fiber ply and matrix

are rather similar ([190,197]) which led to both fiber breakage and matrix

cracking mechanisms to dominate in the transverse impact failure of natural

fiber composite laminates. Post-impact SEM imaging of the fracture surfaces in

Figure 6.12a to d showed that fiber-matrix debonding, fiber fracture, fiber pull

out, and generally brittle fracture of the matrix with low-to-moderate amounts

of matrix yielding in some locations were found to be the major mechanisms

accountable for dissipation of the impact energy.


Chapter 6 159

(a) (b)

A B A
A − Fiber fracture (warp fibers) B − Fiber/matrix debonding (weft fibers) C − Matrix A − Fiber bundle fracture

(c) (d)
B

A
B
A − Fiber fracture B − River marking C − Cleavage marking D −Fiber pulled out A −Matrix crack B − Cleavage marking

Figure 6.12. Scanning electron microscopic (SEM) of fracture surfaces of the woven flax fiber
laminate under transverse impact; (a) warp fiber fracture and weft fiber/matrix debonding at
the fracture surface, (b) fiber bundle fracture, (c) and (d) river marking, cleavage marking and
cracking of matrix.

To assess the effect of stitching on damage development in the woven

composite under transverse non-perforation impact load, two characteristic

fracture parameters were used. Both of these parameters are associated with the

fiber ply (in-plane) damages since very little interlaminar failure was observed

within the woven flax fiber composites (Figure 6.6). Firstly, the values of P1

were normalized by dividing by the thickness cubed of plates, P1 /t 3, in order to

eliminate the effect of thickness variation on elastic bending stiffness. Secondly,

the ratio of absorbed energy of impact (Eabsorbed ) to the total crack surfaces
160 Low-velocity impact performance of stitched flax/epoxy composite laminates

created during impact, which represents the fracture toughness of composite

under impact, was determined by:

𝐸𝑎𝑏𝑠𝑜𝑟𝑏𝑒𝑑
𝐺𝑖𝑚𝑝𝑎𝑐𝑡 =
𝐿×𝑡 (6.1)

where 𝐿 is the total crack length of the transverse and longitudinal through-

thickness fractures observed, based on the average of the crack length measured

from the impact face and back surface of specimens (as seen from Figure 6.1)

and 𝑡 is the specimen thickness (Table 6.2). The crack surface area (L × t) was

approximated from the multiplication of the total crack length by the specimen

thickness since very little evidence of interlaminar cracking was found (see

Figure 6.6). This applies to the un/stitched woven composites only. The area

contribution from the fracture surface roughness was taken to be negligible in

this case.

The normalized parameters for unstitched and stitched laminates are shown in

Figure 6.13. It was observed that both the normalized damage initiation load

(𝑃1 /𝑡 3) and the impact fracture toughness values (𝐺𝑖𝑚𝑝𝑎𝑐𝑡 ) are about 6% and

16% lower for the flax yarn stitched laminates, respectively. On the other hand,

the results of the cotton thread stitched laminates were lower but not statistically

different from the unstitched woven laminates.

This difference may be related to the cross-sectional area of the stitch, where

the flax yarn is six-times bigger than cotton thread as discussed in Chapter 4

[180]. Similar to stitched synthetic fiber composites [129], stitches in natural

fiber composites would also perform as a crack initiator within the composites.
Chapter 6 161

Since the stitch areal fraction is equivalent for both stitched laminates, the

thicker stitches (i.e. flax yarn) appeared to lead to greater decline in the impact

fracture toughness of the laminates. In other words, the stitch-induced defects

associated with the dimension of stitches, such as in-plane reinforcing fiber

distortion and resin-rich pockets, has a direct influence over the impact damage

resistance of stitched natural fiber composites.

40 34

32.8 33
EAbsorbed /Damage area (KJ/m2)

32.2
32

P1 /t3 (N/mm3)
35

31
30.5
31.7
31.1 30

30
29

26.7 28

25 27
PW0 PWC PWF

Woven flax fiber laminate

Figure 6.13. Effect of stitch fiber on damage initiation load and fracture toughness under
impact load.
162 Low-velocity impact performance of stitched flax/epoxy composite laminates

Table 6.3. Summary of impact test results.

Impact Energy P1 Pmax P1/t3 Eabsorbed Eabsorbed/E impact


Sample
(J) (N) (N) (N/mm3) (J) (%)

PW0 2112 ±11 2534 ±29 32.8 ±0.7 13.4 ±0.3 68.2 ±1.5
PWC 1939 ±19 2105 ±12 32.2 ±0.6 17.0 ±0.1 86.5 ±0.5
19.6
PWF 1938 ±12 2142 ±46 30.5 ±0.6 15.8 ±0.1 80.4 ±0.5
UD0 2515 ±32 2935 ±33 40.2 ±0.8 14.7 ±0.1 75.1 ±0.5
PW0 2266 ±48 2661 ±25 33.9 ±0.7 43.1 ±0.4 96.7 ±0.9
PWC 2105 ±64 2701 ±33 32.6 ±0.7 43.2 ±0.3 96.8 ±0.8
44.6
PWF 2093 ±26 2339 ±25 33.2 ±0.7 39.6 ±0.9 88.7 ±2.0
UD0 2618 ±56 3159 ±52 41.2 ±0.8 44.2 ±0.2 99.2 ±0.5

6.7. Conclusion
An experimental investigation of natural fiber stitching on the low-velocity

impact response of continuous flax/epoxy composite laminates has been

presented in this chapter. The flax and cotton stitched woven composites

showed consistent and similar impact response despite differences in their stitch

fiber type and thickness. Fiber fracture was prevalent, hence the impact damage

was mainly in the form of a cross-shaped crack (along the weft and warp

directions) that is visible on both top/bottom surfaces, with very little

delamination. In all samples, stitching was detrimental to the impact resistance

of the composites. The normalized initiation force (P1 /t 3) for stitched woven

flax laminates were lower than the unstitched ones, particularly when flax yarn

stitches were used. This was attributed to the lower crack resistance of the resin-

rich regions that reside between the stitch loops. Flax yarn stitching led to a

reduction of the impact fracture toughness of the woven flax fiber laminate by

almost 16%, while this reduction for the cotton stitching was only 5%.
Chapter 6 163

Therefore, the flax yarn stitching of woven flax laminate appeared to be more

detrimental to the structural performance of the composite under low-velocity

impact. At higher impact energies where the laminates were fully perforated,

the efficiency of absorbed energy over impact energy all exceeded 90%, except

for the flax yarn stitched composite which had the worst performance. Similar

degrees of perimeter damage joining the transverse and longitudinal crack tips

were observed due to bending failures of the quadrantal sections. Finally, for

the unstitched panels, unidirectional cross-ply fiber preform construction was

found to provide greater impact resistance than the stacked woven fiber preform.
164 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

Chapter 7
Numerical analysis of low-velocity impact

on woven flax/epoxy composite laminate

7.1. Introduction
In this chapter, a finite element (FE) modeling of low-velocity impact onto

woven flax fiber composite is presented to study the material behavior and

damage mechanisms for the impact energies experimentally studied in previous

chapters. A continuum damage mechanics (CDM) based approach is applied to

model the in-plane fiber and matrix damage (intralaminar failure) using Abaqus

6.11/Explicit package with a built-in VUMAT, and delamination (interlaminar

failure) is modeled using cohesive elements with a bilinear cohesive law. The

model is validated with the experimental results for both the non-perforation

and perforation impact cases. Finally, the intralaminar and interlaminar damage

evolution and energy absorption mechanisms of the woven flax fiber composite

under impact loading are discussed in detail.


Chapter 7 165

7.2. Modeling strategy


A 3D finite element model is created to analyse the unstitched woven flax fiber

composite with the lay-up of [0]4 subjected to low-velocity impact at two

different energies. The drop weight impact test is simulated in Abaqus

6.11/Explicit according to the experimental set up presented in Section 3.10. A

built-in VUMAT code is used to model the intralaminar damage of the woven

composite and the interlaminar damage is modeled using cohesive elements.

The damage model is based on the smeared crack approach with the maximum

stress failure criterion and elastic damage in the fiber directions and a nonlinear

elastic-plastic damage model for in-plane shear response. The experimental

results of impact onto the woven flax fiber composite studied in Chapter 6 are

used to validate the model. Material properties used in this model are

summarized in Table 7.1.

The 3D model of the composite is 75mm in width, 125mm in length and

4.11mm thick, with the lay-up of [0]4, which is used to capture the impact

damage of the unsupported central region of the test specimen, as shown in

Figure 7.1.
166 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

100 mm
(a)
75 mm

125 mm
150 mm

Modeled region

Actual Specimen

Supported region
(b)
Flax Fabric ply
Ply No. 1
Coh. No. 1 Thickness= 1.02 mm
Ply No. 2
Coh. No. 2
Ply No. 3 Cohesive layer
Coh. No. 3
Ply No. 4 Thickness= 0.01 mm

Figure 7.1. Schematic of (a) the model dimension, and (b) the lay-up used.

The 3D FE model used to analyse low-velocity impact test of the unstitched

woven flax fiber composite is shown in Figure 7.2. The boundary conditions are

introduced by limiting the rotations and the displacement at edges of the plate.

Each of the fabric plies is modeled using quadrilateral continuum shell elements

(SC8R). The interaction between the woven flax fiber plies is modeled using a

single layer of 3D cohesive elements (COH3D8) with a thickness of 0.01 mm.

The nodes of both faces of the cohesive elements are tied to neighboring

structural elements. The impactor is modeled by a hemispherical rigid surface

(R3D3) with a point mass of 9.99 kg, and an initial velocity of 1.893 m/s and

2.926 m/s according to the experimental test data for non-perforation and

perforation impact cases, respectively.


Chapter 7 167

Ply 1
Ply 2
Ply 3
Ply 4

A Woven flax fiber composite test specimen; B Rigid impactor

Figure 7.2. 3D finite element model of the drop-weight impact test.

The model was meshed with 0.75 mm by 0.75 mm element size and two

elements with a thickness of 0.51 mm were used through-the-thickness of each

fabric ply. The element size was determined based on the rule proposed by

Harper [198] for the cohesive zone length to ensure that the element is small

enough to capture the stress gradient at the wake of crack tips. According to this

method, the cohesive element size must be smaller than one-third of the

predicted characteristic length for the cohesive zone. The cohesive zone length

is generally smaller under Mode I loading condition [182]. Therefore, in this

study, the mesh size is chosen based on the cohesive zone length for Mode I

loading. Besides, a mesh convergence study is conducted with 3 different

meshes, 0.75 mm, 0.60 mm, 0.5mm to ensure the mesh-independency of results.

The interaction between the impactor and plate is simulated as a hard contact

using the Abaqus global contact algorithm with a frictionless tangential

behavior. In order to overcome the hourglass issue, the enhanced strain based
168 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

method available in Abaqus mesh control is employed. The elements were set

to be deleted when either the damage parameter in fiber directions, 𝑑1 or 𝑑2

<0.99 was satisfied.

7.3. Material model


The continuum damage model proposed by Johnson [158] is considered for

modeling the material behavior of the plain woven flax fiber composite. The

model is based on the following simplifying assumption:

 The reinforcing fibers are continuous and aligned. In other words, the

effect of complex microstructures produced by the interlacing and

undulation of flax fiber tows are neglected. Therefore, the principle

directions of the material coordinate system are representing the fiber

directions.

 The warp and weft fibers are orthogonal and aligned with the principle

directions of the material coordinate system (plain weave).

 The material response in the fiber directions is linear.

 Fiber and matrix damage modes are decoupled. Fiber damages are only

under tension/compression while matrix damage occurs in shear.

 Fiber properties may be different in tension and compression.

 Model sustains the damage parameters unchanged in unloading (non-

healing). It is only changed if higher damage load is re-applied.

 Plastic strains of the woven composite are only associated with (in-plane)

matrix shear response.


Chapter 7 169

7.3.1. Elastic behavior of woven laminate

The woven fabric reinforced composite ply is modeled as a homogeneous

orthotropic elastic material in which the damage prior to ultimate failure can be

captured by progressive stiffness degradation. As mentioned above, it is

assumed that the warp and weft fibers are orthogonal and aligned with the

principle axis of local coordinate system. Figure 7.3 illustrates a schematic of

2D woven fabric reinforcement and corresponding local material directions

used in the constitutive model.

Weft fibers
2

Warp fibers

Figure 7.3. Schematic of 2D woven fabric reinforcement and the local coordinate system.

A continuum damage mechanics (CDM) approach is used to model the in-plane

fiber and matrix damage (intralaminar failure) in which the elastic properties of

the ply are degraded by damage variables governing by damage evolution

equations [158]. The general constitutive law of elastic orthotropic material for

plane stress is defined by Eq.(7.1)

𝝐 = 𝑺𝝈
(7.1)
170 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

where 𝝐 = (𝜖11 , 𝜖22 , 2𝜖12 )𝑇 and 𝝈 = (𝜎11 , 𝜎22 , 𝜎12 )𝑇 are the in-plane strain

and stress component of composite ply with a symmetry material (fiber)

directions of 1 and 2. The elastic compliance matrix of 𝑺 is given by

1 −𝑣12
0
𝐸1 (1 − 𝑑1 ) 𝐸1
−𝑣21 1
𝑆= 0
𝐸2 𝐸2 (1 − 𝑑2 )
1
0 0
[ 2𝐺12 (1 − 𝑑12 )]
(7.2)

where the elastic constants 𝐸1 and 𝐸2 are the Young’s modulus in the directions

of 1 and 2, 𝐺12 is the in-plane shear modulus and 𝑣12 is the in-plane Poisson’s

ratio. Scalar variables of 0 ≤ 𝑑𝑖 ≤ 1 are the damage parameters and represent

the coefficient of reduction in the elastic modulus of material after micro-

damage initiation. Here, 𝑑1 and 𝑑2 are the fiber damage parameter along the

directions of 1 and 2, respectively, and 𝑑12 is associated with the in-plane shear

damage of matrix. In the current model, for reason of simplicity, there is no

independent damage variable to degrade the Poisson’s ration. Therefore, it can

be shown through the uniaxial stress field that 𝑣12 and 𝑣21 are reduced by the

factor of (1 − 𝑑1 ) and (1 − 𝑑2 ), respectively [158].

In order to model the damage evolution of 2D fabric reinforcement, it is

assumed that in the both fiber direction, fiber failure is dominant under tension

and compression, and matrix failure is dominant under in-plane shear. It is also

assumed that fiber and shear damage modes (under normal and shear stresses)

are decoupled. These simplifying assumptions for the elastic damage mechanics

of fabric composite plies were discussed in detailed by Johnson [158].


Chapter 7 171

7.3.2. Fiber failure

The maximum longitudinal stress failure criterion is used to determine fiber

failure in both principle direction (𝛼 = 1,2)

𝑋𝑓−𝛼 < 𝜎̃𝛼 < 𝑋𝑓+𝛼


(7.3)

where 𝑋𝑓+𝛼 and 𝑋𝑓−𝛼 are the ultimate tensile and compressive strength of the

composite in the fiber directions, respectively. The damage activation functions,

𝐹𝛼 , is used to define the elastic domain at any given time as

𝐹𝛼 = 𝜙𝛼 − 𝑟𝛼 ≤ 0
(7.4)

𝜎
̃
where 𝜙𝛼 = |𝑋𝛼 | defines the failure criteria in which 𝑋𝛼 = 𝑋𝑓+𝛼 or 𝑋𝛼 = 𝑋𝑓+𝛼 ,
𝛼

and 𝑟𝛼 is the damage thresholds (𝛼 = 1,2), which are assumed to comply the

Kuhn-Tucker complementary conditions as

𝐹𝛼 ≤ 0, 𝑟𝛼̇ ≥ 0, 𝑟𝛼̇ 𝐹𝛼 = 0
(7.5)

and the consistency condition of 𝑟𝛼̇ 𝐹𝛼̇ = 0 . The evolution of the damage

variables 𝑑1 and 𝑑2 are monotonic increasing and based on the assumption that

the total energy needed to fail an element has to be equal to the fracture energy

per unit area under uniaxial tensile or compressive loading. The energy of

failing an element is determined from the area under the stress-strain curve,

multiplied by the characteristic length of the element as


∫ 𝜎𝛼 𝑑(𝐿𝑐 𝜀𝛼 ) = 𝐺𝑓𝛼
0
(7.6)
172 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

where 𝐺𝑓𝛼 (𝛼 = ±1, ±2) is the tensile or compressive fracture toughness of the

composite material in the principle fiber directions (or in-plane fracture

toughness), and 𝐿𝑐 is the characteristic length of the element which is defined

as the square root of the area of a continuum shell element in Abaqus [185].

Here, the values of the intralaminar fracture toughness used (Table 7.1) are

estimated from the compact tension test results of the woven flax fiber

composites (with tensile strength and stiffness of 88.39 MPa and 6.88 GPa,

respectively) carried out by Liu et al. [9].

7.3.3. Matrix failure

Deformations of the fabric composite plies under in-plane shear are controlled

by matrix elastic response, plasticity and matrix cracking. Thus, matrix

dominated failure is the main failure mechanism of fabric composite plies under

in-plane shear. In order to model the shear response, the total shear strain in the

𝑒𝑙 𝑝𝑙
ply, 𝜀12 , is split into the sum of elastic, 𝜀12 , and plastic strain, 𝜀12 , as

𝑒𝑙 𝑝𝑙
𝜀12 = 𝜀12 + 𝜀12
(7.7)
Using the constitutive law of elastic response of material, the effective shear

stress, 𝜎̃12 , is calculated as

𝜎12 𝑒𝑙
𝜎̃12 = = 2𝐺12 𝜀12
1 − 𝑑12
(7.8)

where the elastic strain is obtained from Eq.(7.7). Assuming that only the

effective shear stress (𝜎̃12 ) leads to plastic deformation, the shear yield function

takes the form of


Chapter 7 173

𝐹12 = |𝜎̃12 | − 𝜎̃0 (𝜀̅𝑝𝑙 ) ≤ 0


(7.9)

where the hardening function of 𝜎̃0 is defined as

𝜎̃0 (𝜀̅𝑝𝑙 ) = 𝜎̃𝑦0 + 𝐶(𝜀̅𝑝𝑙 )𝐾


(7.10)

in which the hardening parameters of 𝜎̃𝑦0 , 𝐶 and 𝐾 can be obtained by a cyclic

tensile test of a fabric composite loading at 45° respect to the principle fiber

directions [158]. The shear response of the woven flax reinforced composite

tested under monotonic and cyclic loading with six load-unload cycles are given

in Figure 2. The procedure for determining the hardening parameters is given in

Appendix C.

In order to activate the shear damage, the maximum shear stress failure criterion

is used. Therefore, the elastic domain function for shear becomes

𝐹12 = 𝜙12 − 𝑟12 ≤ 0


(7.11)

𝜎
̃12
where 𝜙12 = , the criteria for the shear damage initiation is based on the
𝑆

maximum shear yielding stress of 𝑆 for matrix damage, and 𝑟12 is the (shear)

damage thresholds.
174 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

50 Monotonic loading
Cyclic loading

40
Shear stress, σ12 (MPa)

30

20
-45 ° +45 °

10 (1 − 𝑑12 )𝐺12

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Shear strain, ε12 (%)

Figure 7.4. Monotonic and cyclic shear stress-strain curves obtained from tensile tests on ±45
woven flax/epoxy laminate.

7.3.4. Delamination

Delamination or interlaminar failure of composite laminates is one of the most

critical failure modes which can deteriorate their load carrying capacity.

Transverse impact usually creates intensive interlaminar shear stress due to

bending of plate under point loading. Due to the bending stiffness mismatch,

the interlaminar shear stress increases when the laminate is thick or the

thickness of fiber plies is high. The high interlaminar shear stress leads to

delamination which generally runs in the resin rich area between the fiber plies.

In this chapter, the Abaqus cohesive element (COH3D8) with a bilinear

traction-separation law is used to model the interlaminar interface of the woven

flax composite. It is assumed that delamination occurs in the homogeneous

resin-rich region between plies and the R-curve effect is neglected. Thus, only
Chapter 7 175

the initiation value of Mode II interlaminar fracture toughness are used. The

procedure of delamination modeling is presented in Chapter 5.2.2.2. The Mode

I and II interlaminar fracture toughness of woven flax fiber composites used in

the delamination modeling were measured through DCB and End Notched

Flexure (ENF) tests.

All material properties used in the numerical simulation of woven flax fiber

composite are summarized in Table 7.1.

Table 7.1. Material properties used in numerical modeling.

Elastic properties 𝐸1 = 7.9 GPa; 𝐸2 = 9.5 GPa; 𝐺12 = 3.98 GPa; 𝑣12 = 0.26
𝑋𝑓+1 = 91MPa; 𝑋𝑓−1 = 82† MPa;
In-plane strength
𝑋𝑓+2 = 105 MPa; 𝑋𝑓−2 = 93† MPa; 𝑆 = 21 MPa
Non-linear shear properties 𝜎̃𝑦0 = 23.9 MPa; 𝐶 = 68; 𝐾 = 0.45
𝐺𝑓+1 = 4.1 KJ/m2 ; 𝐺𝑓−1 = 3.5† KJ/m2 ;
In-plane fracture toughness*
𝐺𝑓+2 = 4.7 KJ/m2 ; 𝐺𝑓−2 = 3.7 †KJ/m2
𝐺𝐼𝐶 = 3.2 KJ/m2; 𝐺𝐼𝐼𝐶 = 3.8 KJ/m2 ; 𝜂 = 1;
Interlaminar properties
σCI = 35 Mpa; σCII = 50 Mpa
† Estimated values.
∗ Values adapted from [9].

7.4. Numerical simulation results and discussion


7.4.1. Impact force response

The predicted impact force-time history of unstitched woven flax fiber

composite for non-perforation (19.6J) and perforation (44.6J) impact test are

compared with the experimental results in Figure 7.5. As can be seen, the impact

force prediction is in good agreement with the experimentally measured results,

particularly for non-perforation impact. The model is able to capture the initial
176 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

elastic response, damage initiation and progression and unloading process of the

woven flax fiber laminates with a reasonable accuracy.

In perforation impact (44.6J), the numerical force response up to peak force (𝑇2 )

fits very well with the results from the experiment, albeit with slightly more

intense force disturbances at damage initiation. After peak force (from 𝑇2 to 𝑇4)

however, the force behavior is slightly different than the experimental response.

This discrepancy can be due to the instability occurs during perforation test. It

was observed from the videos captured by the high-speed camera during impact

test that the impactor may sometimes rotate slightly by 1-1.5° due to the special

design of the drop weight apparatus which leave a small clearance in sliding

guides. Consequently, the impactor deviates slightly from the original vertical

alignment while penetrating into the plate. This, together with contact friction

effect which are neglected in the FE modeling, would also contribute to the

deviation of the ideal model set up (which is constrained to move only along the

out of plane direction) from the actual experimental measurements. The very

small rotation of the impactor during perforation impact tests is also suspected

to be related to the unsymmetrical damages on the laminates, as shown in

Figure 6.2.
Chapter 7 177

(a) (b)
3000 Exp-#1 3500 Exp-#1
T1 T2 T3 Exp-#2 T1 T2 T3 Exp-#2
Exp-#3 Exp-#3
FE simulation 3000 FE simulation
2500

2500
T4 T4
2000

2000

Force (N)
Force (N)

1500

1500

1000
1000

500
500

0 0
0 5 10 15 20 25 0 5 10 15 20 25
Time (ms) Time (ms)

Figure 7.5. Impact force–time curves of numerical simulation versus experimental results for
(a) 19.6J (non-perforation) and (b) 44.6J (perforation).

Figure 7.6 compares the load-displacement curves obtained from the numerical

simulation with the experimental results for both the impact energies. In the case

of non-perforation impact, the numerical results show excellent agreement with

the experimentally measured force-displacement. The all three zone (i.e. A –

elastic response, B – damage development and C – unloading) for the non-

perforation impact discussed in Chapter 6.5, were captured with good accuracy.

The FE simulation for the perforation impact in Figure 7.6b again matches well

initially (up to 𝑇2 ) but deviated from the experimental data in the later part of

the test. It is observed that while the experimental loads drop off at 𝑇3 which

coincides with the fracture of a failing quadrant, the FEM did not predict the

quadrant failure and simulated a gradually declining load that would be

characteristic of a structure which had remained intact. This is affirmed later in

the damage evolution data presented in Section 7.4.2, Figure 7.7b, that at 𝑇3 and

𝑇4, the quadrants had remained largely intact in the FE model.


178 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

(a) (b)
3000 3500
T2 Exp-#1
T2
T1 Exp-#2
T3
3000 Exp-#3
2500 T1
T3 FE simulation

2500
2000

2000

Force (N)
Force (N)

1500

1500

1000
1000

Exp-#1
500 T4 T4
Exp-#2 500
Exp-#3
FE simulation
0 0
0 2 4 6 8 10 12 0 5 10 15 20 25 30 35

Displacement (mm) Displacement (mm)

Figure 7.6. Impact force–displacement curves of numerical simulation versus experimental


results for (a) 19.6J (non-perforation) and (b) 44.6J (perforation).

7.4.2. Impact damage evolution

The damage evolution of both the impact cases associated with the critical time

𝑇1 to 𝑇4 of impact event, illustrated on the force responses of Figure 7.5 and

Figure 7.6, are presented here. Figure 7.7a and b show the damage contours on

the impact and back face of the woven flax fiber laminate as well as all cohesive

interfaces at the specified time 𝑇1 to 𝑇4. Based on the force responses and the

corresponding damage contours, the damage evolution for each of the impact

cases can be described as below:

7.4.2.1. 19.6J (non-perforation)

 The interval from 𝑇𝑖𝑚𝑒 = 0 up to 𝑇1 represents the elastic response of

the specimen under bending until the force of yielding is reached and

the initial signs of material damages are appeared at 𝑇𝑖𝑚𝑒 ≅ 𝑇1 . The

fiber and matrix damages are first formed on the fiber plies prior to any

failure in the cohesive interface (delamination). Then, shortly after the


Chapter 7 179

in-plane damages initiation, the first sign of interface failure is appeared

on the most top cohesive interface (i.e. Coh.1) under the impact place.

 Starting from 𝑇1 , the in-plane damages develop along the fiber

directions (weft and warp) at the centerlines, forming the cross shaped

crack in the plate. At this stage, the impact force still increases at a lower

rate compared to initial elastic response, while the velocity of impactor

decrease. At 𝑇𝑖𝑚𝑒 = 𝑇2 , the force reached to its maximum value. The

damage of cohesive interfaces also follows the in-plane damage pattern

and it is almost limited into a band with a width equivalent to the in-

plane damage width.

 From 𝑇𝑖𝑚𝑒 = 𝑇2 to 𝑇3 , the damages still propagate, however, the

propagation rate is much lower as the majority of impact energy has

been transferred to the laminate in the form of elastic bending and

damage mechanism.

 At 𝑇𝑖𝑚𝑒 = 𝑇3 , the impactor fully stops (motion reversal point in

Figure 7.6a) and the laminate reaches to a maximum deformation.

Damage development in the laminate stops size and the maximum

damage size is achieved is at this point.

 From 𝑇𝑖𝑚𝑒 = 𝑇3, the unloading process is begun and the elastic energy

stored in the laminate is delivered to the impactor while no further

damage is formed in the laminate as the damage contours remain

unchanged up to 𝑇4 (end of the test).

The final damage contours on the both faces are also compared with the actual

specimens in Figure 7.7a. At the impact face, the relative errors between the
180 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

predicted and experimental length of the longitudinal and transverse fiber-

dominated cracks (as defined in Figure 6.1) are 2% and 8%, respectively. This

error at back surface is 2% for transverse crack and 16% for longitudinal crack.

The underestimated crack length in the longitudinal direction can be associated

with the fully clamped boundary conditions used in the numerical model, while

the rubber tip toggle clamp used to fix the test specimen at four corners (see

Figure 3.13) may not offer an ideal clamped boundary condition. Nevertheless,

the model has the capacity of predicting the material behavior of the woven flax

composite subjected to low-velocity impact with a satisfactory accuracy.

7.4.2.2. 44.6J (Perforation)

 The impact event up to 𝑇2 is pretty similar to what described for the

19.6J case, however, the duration is shorter obviously because of the

higher impact velocity in the 44.6J impact case. Moreover, unlike the

impact of 19.6J, fiber damage forms at the clamped edges (particularly

along the long edges) due to the higher impact load.

 From 𝑇𝑖𝑚𝑒 = 𝑇2 to 𝑇4 both the intralaminar and interlaminar damages

propagates continuously while the impactor is moving further into the

specimens. At 𝑇𝑖𝑚𝑒 = 𝑇3 , the damage starts to deviate from the original

pass because of the bending failures of the quadrants created. Finally, a

circular shaped damage is formed due to bending failure of the quadrants.

 At 𝑇𝑖𝑚𝑒 = 𝑇4, the impactor passes through the plate and the damage

evolution is stopped.

The final fiber and matrix damage zone on top (Ply 1) and back (Ply 4) faces

are in good agreement with the visual damage size on the actual impacted
Chapter 7 181

specimen. As seen from Figure 7.7a and b, the damage contour of cohesive

interfaces in both the impact cases follows the in-plane damage pattern and is

almost limited into a band with a width of a few elements. This is associated

with the relatively low in-plane strength of the impregnated woven flax fiber

lamina as well as the high interlaminar toughness of this laminated composite.

These characteristics of this composite system lead to in-plane fiber dominated

fractures prior to the interlaminar damage development under the stress

distribution caused by transverse impact loading of the laminate. This result is

very similar to what observed experimentally from the impacted specimen

(Figure 6.6). The very small interlaminar failure zone (cohesive interface)

which observed from both the experimental results and the numerical prediction,

indicates that delamination is not a dominant failure mode in the woven

flax/epoxy laminated composite.


182 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

(a)
75mm
Ply 1 Ply 4
Experimental

Ply 1

125mm
Coh. 1
Coh. 2
Coh. 3
Warp Warp
Ply 4
Weft Weft
T4
T3
T2
T1

Ply 1 Coh. 1 Coh. 2 Coh. 2 Ply 4

Impact face Cohesive interfaces Back face


Chapter 7 183

(b)
75mm
Ply 1 Ply 4
Experimental

Ply 1

125mm
Coh. 1
Coh. 2
Coh. 3
Warp Warp
Ply 4
Weft Weft
T4
T3
T2
T1

Ply 1 Coh. 1 Coh. 2 Coh. 2 Ply 4

Impact face Cohesive interfaces Back face

Figure 7.7. Damage evolution at top and back surface of composite plate and the all cohesive
interfaces during low-velocity impact at specified time T 1 to T4 for (a) 19.6 J and (b) 44.6J
impact energy (The damage at T 4 is compared with the actual specimen).
184 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

7.4.3. Matrix damage and delamination

In-plane matrix damage was found to absorb substantial energy compared to

delamination under transverse impact loading. Figure 7.8 indicates that the

energy dissipated by matrix damage during impact is significantly higher than

the energy dissipated by delamination for both impact cases. The energy

dissipated by delamination is 3% and 4% of total absorbed impact energy of

non-perforation and perforation test, respectively, whereas the contribution of

matrix damage is almost 55% and 45% for non-perforation and perforation

impact, respectively. The rest of absorbed energy is dissipated by fiber damage

mechanisms. In other words, fiber breakage and matrix damage composed of

plasticity and yielding are the main energy dissipation mechanisms. These

results of numerical analysis are consistent with the experimental finding of

Chapter 6.6, in which, in the absence of delamination, the fiber breakage and

matrix cracking were found to be the major energy dissipation mechanisms.

The tendency to propagate in-plane damages rather than delamination is

attributable to, first, the high interlaminar fracture toughness, and second, the

low in-plane strength of the woven flax fiber composite, as discussed in Chapter

4. In other words, the energy required for delamination propagation is higher

than the energy needed for developing in-plane damages in the woven flax

laminate under transverse impact. Consequently, the majority of the impact

energy is dissipated by the in-plane damage mechanisms which are easier to

initiate and propagate. This behavior is in contrast to that generally observed for

the laminates made of synthetic fibers where delamination is the dominant


Chapter 7 185

failure mode under impact loading due to a significant difference between the

in-plane interlaminar strength [129,199].

(a) (b)
10 25
Delamination Matrix damage Delamination Matrix damage
9

8 20

6 15

Energy (J)
Energy (J)

4 10

2 5

0 0
0 5 10 15 20 25 0 5 10 15 20

Time (ms) Time (ms)

Figure 7.8. Comparison of energy dissipation by delamination with matrix damage in woven
flax fiber composite subjected to (a) 19.6J and (b) 44.6J impact energy.

7.4.4. Permanent deformation

The permanent out-of-plane deformation of the plate is mainly attributed to the

plasticity of matrix [84]. Figure 7.9 shows the permanent indentation captured

by numerical simulation, after full separation of the impactor in non-perforation

impact (19.6J), and compares with an experimental specimen. This permanent

deformation is determinative in defining the post-impact performance of the

laminates [200].

The prediction of the permanent indentation after impact ( = 3.42 mm) is in good

agreement with the experimental values ( = 3.95 ∓ 0.15 mm ) which are

measured immediately after test. The capability of the numerical model to

capture the permanent deformation is attributed to the nonlinear shear response

of the matrix damage model. As discussed in the martial modeling, the matrix
186 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

elastic and inelastic parameters were empirically calibrated through the cyclic

shear test (see Appendix C).

(a) (b)

A A

Permanent indentation A A

Figure 7.9. Permanent indentation after non-perforation impact (19.6J); (a) average
experimental measurements = 3.95 mm; (b) numerical results = 3.42 mm.

The 3D deformed shape of the perforated specimen (44.6J) captured by this

model is given in Figure 7.10. It can be seen that the overall damage pattern,

damage size, and permanent deformation after penetration are in good

agreement with the experimental impacted specimen of Figure 6.2a.

Impact face Back face

Figure 7.10. Deformed shape after full perforation impact (44.6J)


Chapter 7 187

7.4.5. Effect of intralaminar fracture toughness

It was shown in Chapter 6.6 that the stitching caused imperfections can play as

a crack initiator which result in a lower intralaminar toughness of the composite

when subjected to transverse impact. This effect of stitching is investigated

using the developed numerical model while the values of used intralaminar

fracture toughness for unstitched laminates (as given in Table 7.1), are reduced

by 16%. This value is used based on the maximum reduction in the impact

fracture toughness of the laminate which was empirically observed for the flax

yarn stitched woven laminates (PWF) – as shown in Figure 6.13. It is

worthwhile to note that here only the effect of reduction in intralaminar fracture

toughness of the laminate is evaluated and the other properties remain

unchanged.

The results of numerical simulation for the reduced intralaminar fracture

toughness of the woven flax fiber composite (denoted by ‘FEM–reduced FT’)

are compared with the experimental result of the flax yarn stitched laminate

(PWF) in Figure 7.11a. As can be seen, the numerical model with reduced

toughness gives a satisfactory estimation of the impact force for the stitched

laminate. The prediction of the enlarged fiber-dominated cracks at impact and

back surfaces are also in good agreement with those of experimental

measurements, as shown in Figure 7.11b. From the experimental and numerical

results, it can be concluded that although the flax yarn stitching was found to be

effective in enhancing the interlaminar fracture toughness of the flax fiber

composites [180], the intralaminar fracture toughness of the laminates may be


188 Numerical analysis of low-velocity impact on woven flax/epoxy composite laminate

compromised by stitching-induced defects which results a lower impact

resistance (i.e. larger longitudinal and transverse crack).

(a) (b)
75mm 75mm

3000 Impact face Back face


Exp - Unstitched (PW0)
FEM - Unstitched
2500
Exp - Stitched (PWF)

125mm
FEM - Reduced FT

2000
Force (N)

Warp Warp

Weft
1500 Weft

1000

125mm
500

0
0 5 10 15 20 25 30
Ply 1 Ply 4
Time (ms)

Figure 7.11. (a) Impact force history of unstitched and flax yarn stitched laminates; (b)
Prediction of damage on the impact and back surface of the flax yarn stitched laminate using
reduced FT values.

7.5. Conclusion
A numerical model is developed to study the material behavior and damage

development of the woven flax fiber/epoxy composite subjected to low-velocity

impact with two different energies, with and without perforation. The virtual

low-velocity impact test was implemented by using a continuum damage

mechanics (CDM) approach to model the in-plane fiber and matrix damage

(intra-laminar failure) and the cohesive element to model delamination of the

woven flax laminate, and a rigid body to model the impactor. The numerical
Chapter 7 189

results were verified by the experimental study conducted in Chapter 6.

Although the fiber architecture of the flax/epoxy composite differs from

conventional fiber reinforced composites – in terms of the variability in fiber

length, diameter, alignment, and dispersion – it was shown that the CDM

approach with linear approximation for elastic response is able to predict the

force history, deformation and the impact damage in woven flax composite to

good accuracy at different energy levels. The experimental observation that

delamination is very minor and absorbs only up to 4% of the total absorbed

impact energy was also numerically verified. Moreover, from the damage

evolution and energy absorption mechanisms observed, the simulation results

showed that fiber damage and matrix plasticity are the dominant failure modes

which correspond well to experimental observations.


190 General Conclusion and recommendations

Chapter 8
General Conclusion and recommendations

8.1. General Conclusion


This research aims to study the effects of introducing out-of-plane natural fiber

stitches on the in-plane mechanical properties, the interlaminar fracture

toughness and the low-velocity impact performance of resin-infused

continuous-flax/epoxy composite laminates. The investigations sought to

understand the influence of stitch fiber type, stitch areal fraction and preform

architecture on the failure mechanisms and behaviors. Another objective of this

work is to develop finite element models for natural fiber composites, and

enable the accurate prediction of composite failure under Mode I opening and

transverse impact modes.

Experimental tests showed that the presence of stitches led to notable reductions

in the in-plane tensile strength and stiffness of the stitched woven composites

by as much as 18% and 12% respectively at the highest stitch areal fraction of

0.02, regardless of the type of stitch material used. The flexural modulus of the

stitched composite remained relatively unchanged, whilst the flexural strength

of the composites showed a dependence on stitch material type, registering a

reduction of up to 12% when the thicker diameter flax yarn stitch (0.33 mm2)
Chapter 8 191

was used, and negligible effect with the 0.055 mm2 cotton thread stitching, at

similar stitch areal fractions.

Despite having a greater detrimental effect on the composites’ in-plane

properties, stitching with flax yarn was found to be more effective than cotton

threads in enhancing the Mode I fracture resistance of woven flax/epoxy double

cantilever beams at similar stitch areal fractions. The Mode I crack propagated

in a stable progressive manner through cotton-thread stitched laminate at very

similar fracture energies as the unstitched laminates. The flax-yarn stitched

laminates, however, could increase the fracture energies by up to 21%. A change

in crack growth behavior from stable propagation to stick-slip unstable fracture

was observed, but the average fracture loads and GIC values (including the

reduced values at crack arrest points) were always higher than the average

measurements obtained for the unstitched laminates. These results were

recurrent in flax-yarn stitched composites of having either the unidirectional or

woven in-plane flax reinforcements. The increased toughness was mainly

attributed to (i) the increased fiber bridging, (ii) higher tensile bridging traction

supported by the impregnated flax stitch-yarns and (iii) the increase in matrix

shear yielding around the stitch sites.

An interesting observation was made with regards to the in-plane preform

architecture of unstitched flax composites during DCB Mode I testing. During

the early phases of crack initiation and propagation, the G IC values of the

unstitched woven composites showed a more pronounced R-curve development

than the unidirectional flax preforms. The toughening effect came from higher
192 General Conclusion and recommendations

levels of fiber nesting and fiber bridging, which led to greater energy dissipation

supported by the woven preform during fracture.

The experimental investigation of natural fiber stitching on the low-velocity

impact response of continuous flax/epoxy composite laminates showed that

both flax and cotton stitched woven composites failed in a similar manner,

predominated by a cross-shaped, through-thickness crack. Little signs of

delamination were observed, except in the cross-ply unidirectional composite

specimens. The normalized initiation force ( P1 /t 3 ) for stitched woven flax

laminates were marginally lower than the unstitched ones due to lowered crack

resistance brought about by stitch-induced damages. The flax yarn stitch is

generally more detrimental to the transverse impact fracture resistance of the

woven flax composites, reducing the impact fracture toughness by almost 16%,

compared to 5% for the cotton stitch. At the higher impact energies where the

laminates were fully perforated, the efficiency of absorbed energy over impact

energy exceeded 90% for all types of specimens, with a less pronounced

discrepancy between unstitched and stitched laminates. Bending fractures of the

quad-segments created by the cross-shaped cracks were a common feature.

The stitch-enhanced delamination resistance observed earlier in Mode I for flax-

stitched composites could not be replicated in the transverse impact tests, due

to the predominance of fiber failures which severely limited the occurrence of

delamination. The transverse impact damage mode is therefore very different

from those of synthetic carbon or glass fiber composites where the fibers rarely

fail at low loads and most of the damage occurs in the matrix via sub-surface
Chapter 8 193

delamination. The presence of non-visible damage is therefore less likely to be

an issue on the maintenance of natural fiber composites.

Two finite element numerical models were also developed in this project to

provide validated simulation tools for the accurate prediction of natural fiber

composites failure under DCB Mode I opening and transverse impact modes. In

the virtual DCB test model, a cohesive element with bilinear softening law

obtained from superposition of two linear softening cohesive law was used to

model the R-curve effect. The parameter definitions were based on the

experimental Mode I interlaminar fracture toughness values of the UD

flax/epoxy composite laminates. This was superposed with transverse beam

elements as the stitch fibers, based on impregnated stitch yarn properties, to

simulate the delamination of the stitched-laminate in the presence of extensive

in-plane fiber bridging. This FE modeling strategy was able to predict the

change in failure mode to stick-slip propagation in stitched flax composites, and

the Mode I critical strain energy release rates (GIC) obtained from the

delamination simulation study aligned very well with the range of the measured

values. A parametric study on the strength of the stitch yarn showed that the

variation in the measured values of GIC mainly attributes to the intrinsic

dispersion of the stitch yarn strength.

Finally, an FE model was also implemented to study the material behavior and

damage development of the woven flax fiber/epoxy composite subjected to low-

velocity impact. The virtual drop weight impact test was implemented by using

a continuum damage mechanics (CDM) approach to model the in-plane fiber

and matrix damage (intra-laminar failure) and the cohesive element to model
194 General Conclusion and recommendations

delamination of the woven flax laminate, and a rigid body to model the impactor.

The numerical results were verified by the experimental study conducted

in Chapter 6. Although the fiber architecture of the flax/epoxy composite differs

from conventional fiber reinforced composites – in terms of the variability in

fiber length, diameter, alignment, and dispersion – it was shown that the CDM

approach with linear approximation for elastic response is able to predict the

force history, deformation and the impact damage in woven flax composite to

good accuracy at different energy levels. The experimental observation that

delamination is very minor was numerically verified and proved that

delamination failure absorbs only up to 4% of the total absorbed impact energy.

Moreover, from the damage evolution and energy absorption mechanisms

observed, the simulation results showed that fiber damage and matrix plasticity

are the dominant failure modes which correspond well to experimental

observations.

The experiment and numerical work conducted in this study have helped to

deepen the understanding of the material behavior and failure mechanisms of

stitched flax fiber composite laminates. Stitching is certainly an effective

technique to increase the Mode I delamination resistance of natural fiber

composites, but in some measure to the detriment of its transverse impact

performance and in-plane composite properties, and requires careful

consideration to be given to the stitch material and areal fraction. Effective

simulation strategies have been proposed and validated for the numerical

prediction of Mode I delamination and low energy transverse impact fractures,

which can be used for further developments in natural fiber composites.


Chapter 8 195

8.2. Recommendations
The general aims of the research and development work on natural fiber

composites revolves around, firstly, the understanding their material behavior,

fracture mechanism, in-service performance and durability, and secondly, to

improve their properties through new techniques that may extend their use to

many more applications, particularly those that still rely on nonrenewable or

non-eco-friendly materials. Some potential areas for future investigations that

are in line with the aforementioned objectives are proposed below.

 The results of this study suggested that flax-yarn-stitching is a promising

technique to enhance the Mode I interlaminar fracture toughness of

flax/epoxy composite laminates. However, the defects induced by the

stitching process might affect the in-plane performance – i.e. tensile,

flexural and impact – of the woven laminates. On the other hand, it is

known that 3D woven fabric composites generally suggest higher

interlaminar and in-plane fracture toughness, better damage resistance,

and higher energy dissipation in comparison with 2D reinforced

composites [201–203]. Thus, in order to reduce the adverse effects of

stitching, it will be interesting to consider 3D woven textile preforms

using the optimally twisted flax yarn and investigate their mechanical

and impact performance.

 In spite of the favorable properties of flax fibers, their water absorption

and relatively low mechanical properties and impact resistant compared

to synthetic counterparts, are some of the major impediments to the

penetration of their use into higher performance applications. Although


196 General Conclusion and recommendations

numerous studies have attempted addressing these issues in an effort to

enhance the performance of flax and other natural fiber composites

through improving fiber-matrix adhesion [204,205], processing

technique [206], fiber preform architecture [6], their performance is still

considered low [207] for structural applications. The idea of fiber

hybridization, i.e. having two or more types of fiber within the same

matrix, has been around for some time [208], however recent emergence

of new materials creates new and exciting possibilities for obtaining

superior hybrid composites tailored for particular applications.

Motivated by the need for higher mechanical performance, lower cost

and sustainability, there has been growing interest recently in

hybridization of natural fibers with synthetic fibers such as glass [209–

214] and carbon [215–219] with the different objectives vary from

assessing improvements in thermal stability, water absorption behavior,

stiffness, strength and impact properties, to acoustic and vibration

damping properties. However, a comprehensive understanding of the

effect of different hybridization level on the mechanical performance is

still lacking. Future research should attempt to investigate the effects of

the hybridization levels of flax fibers with synthetic carbon or glass

fibers – such as interlayer, intralayer and intrayarn [220]– and quantify

the achievable performance enhancement. Looking at other aspects of

hybridization such as benefits in damage resistance, fatigue behavior,

environmental impacts and recyclability is also necessary in this regard.


Chapter 8 197

 It is also recommended to investigate the possibilities of hybridizing flax

fibers and recycled carbon fibers for manufacturing hybrid aligned

discontinuous fiber composites, and study their performance for

engineering applications.

 In the present work, the potential of the continuum damage model (CDM)

for modeling woven flax fabric composite at ply level was assessed.

Despite the fact that the model (using the average mechanical properties

of the composite) was able to predict the material behavior and damage

development with good accuracy, the intrinsic variation in the size,

distribution, properties and alignment of flax fibers was not taken into

account. One possible avenue of future work is conducting a multi-scale

numerical analysis on the flax fiber composite, starting from the

micromechanics. It would be interesting to numerically study the

progressive failure of flax fiber composite at different length scales,

where it is difficult or impossible to be captured by experimental

methods. To develop a multiscale model, firstly, a representative volume

element (RVE) of flax fiber composite needs to be defined, taking into

account the intrinsic variation in geometry, orientation and mechanical

properties of flax fibers. Then, the results from analysis of this RVE can

be used to determine the meso and microscale behavior of the flax fiber

composites through multiscale analysis approaches. Additional

suggestion would be to develop a model using the statistical mechanics

for predicting the mechanical properties of the composite based on

uncertainty in the properties of fibers. The disposal of numerical models


198 General Conclusion and recommendations

with reasonable accuracy to predict the mechanical properties not only

provide a better understanding of the behavior of natural fiber

composites, but also can help to extend their use to engineering

applications.

 Throughout the study, it was found that flax-yarn-stitching is an

effective way to increase the Mode I fracture toughness of flax fiber

composites. This suggests that flax yarn stitched laminates may be

utilized in applications that are needed for absorbing and dissipating

energy, such as crash beams used in transport vehicles [221]. In this

regards, it would be interesting to study the potential of using stitched

flax fiber laminates for designing lightweight crash structures. Future

research should attempt to investigate the energy absorption

performance and failure mechanisms of these structures under impact

loading.

 In the aspect of low-cost manufacturing, there is a need to conduct a

comprehensive study to assess the effect of various processing

techniques – such as hand lay-up, RTM, VARTM, vacuum infusion, etc.

– of flax fiber composites on the final properties of their composites and

distinguish the key influencing factors. The results of such studies can

help to develop cost-effective products of flax fiber composites and

spread their market.


References 199

References

[1] O. Faruk, A.K. Bledzki, H.-P. Fink, M. Sain, Biocomposites reinforced


with natural fibers: 2000–2010, Progress in Polymer Science. 37 (2012)
1552–1596. doi:10.1016/j.progpolymsci.2012.04.003.

[2] L. Yan, N. Chouw, K. Jayaraman, Flax fibre and its composites – A


review, Composites Part B: Engineering. 56 (2014) 296–317.
doi:10.1016/j.compositesb.2013.08.014.

[3] F. Ahmad, H.S. Choi, M.K. Park, A Review: Natural Fiber Composites
Selection in View of Mechanical, Light Weight, and Economic
Properties, Macromol. Mater. Eng. 300 (2015) 10–24.
doi:10.1002/mame.201400089.

[4] L. Pil, F. Bensadoun, J. Pariset, I. Verpoest, Why are designers


fascinated by flax and hemp fibre composites?, Composites Part A:
Applied Science and Manufacturing. 83 (2016) 193–205.
doi:10.1016/j.compositesa.2015.11.004.

[5] D.U. Shah, Natural fibre composites: Comprehensive Ashby-type


materials selection charts, Materials & Design. 62 (2014) 21–31.
doi:10.1016/j.matdes.2014.05.002.

[6] S. Goutianos, T. Peijs, B. Nystrom, M. Skrifvars, Development of Flax


Fibre based Textile Reinforcements for Composite Applications, Appl
Compos Mater. 13 (2006) 199–215.

[7] B.A. Muralidhar, Tensile and compressive properties of flax-plain


weave preform reinforced epoxy composites, Journal of Reinforced
Plastics and Composites. 32 (2013) 207–213.

[8] F. Bensadoun, K.A.M. Vallons, L.B. Lessard, I. Verpoest, A.W. Van


Vuure, Fatigue behaviour assessment of flax–epoxy composites,
Composites Part A: Applied Science and Manufacturing. 82 (2016)
253–266. doi:10.1016/j.compositesa.2015.11.003.

[9] Q. Liu, M. Hughes, The fracture behaviour and toughness of woven flax
fibre reinforced epoxy composites, Composites Part A: Applied Science
and Manufacturing. 39 (2008) 1644–1652.

[10] C. Scarponi, 12 - Impact behavior of natural fiber composite laminates,


in: N.E. Zafeiropoulos (Ed.), Interface Engineering of Natural Fibre
Composites for Maximum Performance, Woodhead Publishing, 2011:
pp. 341–382.
200 References

[11] S. Liang, L. Guillaumat, P.-B. Gning, Impact behaviour of flax/epoxy


composite plates, International Journal of Impact Engineering. 80
(2015) 56–64.

[12] R. Petrucci, C. Santulli, D. Puglia, E. Nisini, F. Sarasini, J. Tirillò, L.


Torre, G. Minak, J.M. Kenny, Impact and post-impact damage
characterisation of hybrid composite laminates based on basalt fibres in
combination with flax, hemp and glass fibres manufactured by vacuum
infusion, Composites Part B: Engineering. 69 (2015) 507–515.

[13] P. Robinson, E. Greenhalgh, S. Pinho, Failure Mechanisms in Polymer


Matrix Composites: Criteria, Testing and Industrial Applications,
Woodhead Publishing, Elsevier, 2012.

[14] K.T. Tan, N. Watanabe, Y. Iwahori, Impact Damage Resistance,


Response, and Mechanisms of Laminated Composites Reinforced by
Through-Thickness Stitching, International Journal of Damage
Mechanics. 21 (2012) 51–80. doi:10.1177/1056789510397070.

[15] T. Huber, S. Bickerton, J. Müssig, S. Pang, M.P. Staiger, Flexural and


impact properties of all-cellulose composite laminates, Composites
Science and Technology. 88 (2013) 92–98.

[16] S.-C. Dai, W. Yan, H.-Y. Liu, Y.-W. Mai, Experimental study on z-pin
bridging law by pullout test, Composites Science and Technology. 64
(2004) 2451–2457.

[17] Y. Qiao, Fracture toughness of composite materials reinforced by


debondable particulates, Scripta Materialia. 49 (2003) 491–496.

[18] N. Sela, O. Ishai, Interlaminar fracture toughness and toughening of


laminated composite materials: a review, Composites. 20 (1989) 423–
435.

[19] M. Ansar, W. Xinwei, Z. Chouwei, Modeling strategies of 3D woven


composites: A review, Composite Structures. 93 (2011) 1947–1963.

[20] P. Suppakul, S. Bandyopadhyay, The effect of weave pattern on the


mode-I interlaminar fracture energy of E-glass/vinyl ester composites,
Composites Science and Technology. 62 (2002) 709–717.

[21] L. Tong, A.P. Mouritz, M.K. Bannister, 3D Fibre Reinforced Polymer


Composites, Elsevier Science, Oxford, 2002.

[22] Y. Iwahori, T. Ishikawa, Y. Hayashi, N. Watanabe, Study of


interlaminar fracture toughness improvement on stitched CFRP
laminates, Journal of the Japan Society for Composite Materials(Japan).
26 (2000) 90–100.
References 201

[23] Y. Iwahori, K. Nakane, N. Watanabe, DCB test simulation of stitched


CFRP laminates using interlaminar tension test results, Composites
Science and Technology. 69 (2009) 2315–2322.

[24] K.A. Dransfield, L.K. Jain, Y.-W. Mai, On the effects of stitching in
CFRPs—I. mode I delamination toughness, Composites Science and
Technology. 58 (1998) 815–827.

[25] A.P. Mouritz, B.N. Cox, A mechanistic approach to the properties of


stitched laminates, Composites Part A: Applied Science and
Manufacturing. 31 (2000) 1–27. doi:10.1016/S1359-835X(99)00056-1.

[26] K.T. Tan, N. Watanabe, Y. Iwahori, T. Ishikawa, Understanding


effectiveness of stitching in suppression of impact damage: An
empirical delamination reduction trend for stitched composites,
Composites Part A: Applied Science and Manufacturing. 43 (2012)
823–832. doi:10.1016/j.compositesa.2011.12.022.

[27] K.T. Tan, N. Watanabe, M. Sano, Y. Iwahori, H. Hoshi, Interlaminar


Fracture Toughness of Vectran-stitched Composites - Experimental and
Computational Analysis, Journal of Composite Materials. 44 (2010)
3203–3229.

[28] A. Yudhanto, N. Watanabe, Y. Iwahori, H. Hoshi, The effects of stitch


orientation on the tensile and open hole tension properties of
carbon/epoxy plain weave laminates, Materials & Design. 35 (2012)
563–571. doi:10.1016/j.matdes.2011.09.013.

[29] H. Ghasemnejad, Y. Argentiero, T.A. Tez, P.E. Barrington, Impact


damage response of natural stitched single lap-joint in composite
structures, Materials & Design. 51 (2013) 552–560.
doi:10.1016/j.matdes.2013.04.059.

[30] A. Aktaş, M. Aktaş, F. Turan, Impact and post impact (CAI) behavior
of stitched woven–knit hybrid composites, Composite Structures. 116
(2014) 243–253. doi:10.1016/j.compstruct.2014.05.024.

[31] K.T. Tan, A. Yoshimura, N. Watanabe, Y. Iwahori, T. Ishikawa, Effect


of stitch density and stitch thread thickness on damage progression and
failure characteristics of stitched composites under out-of-plane loading,
Composites Science and Technology. 74 (2013) 194–204.
doi:10.1016/j.compscitech.2012.11.001.

[32] A.P. Mouritz, B.N. Cox, A mechanistic interpretation of the


comparative in-plane mechanical properties of 3D woven, stitched and
pinned composites, Composites Part A: Applied Science and
Manufacturing. 41 (2010) 709–728.
doi:10.1016/j.compositesa.2010.02.001.
202 References

[33] X. Sun, L. Tong, M.D.K. Wood, Y.-W. Mai, Effect of stitch distribution
on mode I delamination toughness of laminated DCB specimens,
Composites Science and Technology. 64 (2004) 967–981.
doi:10.1016/j.compscitech.2003.07.004.

[34] G.-C. Tsai, J.-W. Chen, Effect of stitching on Mode I strain energy
release rate, Composite Structures. 69 (2005) 1–9.
doi:10.1016/j.compstruct.2004.02.009.

[35] M.Z. Rong, M.Q. Zhang, Y. Liu, Z.W. Zhang, G.C. Yang, H.M. Zeng,
Effect of Stitching on In-Plane and Interlaminar Properties of
Sisal/Epoxy Laminates, Journal of Composite Materials. 36 (2002)
1505–1526.

[36] A. Hodzic, R. Shanks, Natural Fibre Composites: Materials, Processes


and Properties, Woodhead Publishing, 2014.

[37] P. Wambua, J. Ivens, I. Verpoest, Natural fibres: can they replace glass
in fibre reinforced plastics?, Composites Science and Technology. 63
(2003) 1259–1264.

[38] I. Van de Weyenberg, Flax Fibres as a Reinforcement for Epoxy


Composites (Vlasvezels als versterkingsmateriaal in epoxy
composieten), Ph.D., KU Leuven, 2005.
https://lirias.kuleuven.be/handle/1979/164 (accessed May 7, 2014).

[39] K. Pickering, Properties and Performance of Natural-Fibre Composites,


Elsevier, 2008.

[40] J.M. Lawther, HANS LILHOLT, Comprehensive Composite Materials:


Fiber Reinforcements and General Theory of Composites. 1 (2000) 303.

[41] D.U. Shah, Developing plant fibre composites for structural


applications by optimising composite parameters: a critical review, J
Mater Sci. 48 (2013) 6083–6107. doi:10.1007/s10853-013-7458-7.

[42] Lucintel Analyzes Natural Fiber Composites Outlook for 2011 and
Beyond, (2011). http://www.businesswire.com/.

[43] Carus M, Eder A, Dammer L, Korte H, Scholz L, Essel R, Breitmayer


E, Wood-Plastic Composites (WPC) and Natural Fibre Composites
(NFC): European and Global Markets 2012 and Future Trends in
Automotive and Construction, 2014.

[44] J. Müssig, K. Haag, 2 - the Use of Flax Fibres as Reinforcements in


Composites, in: O. Faruk, M. Sain (Eds.), Biofiber Reinforcements in
Composite Materials, Woodhead Publishing, 2015: pp. 35–85.

[45] H.L. Bos, The potential of flax fibres as reinforcement for composite
materials, Technische Universiteit Eindhoven, Eindhoven, 2004.
References 203

[46] D.U. Shah, P.J. Schubel, M.J. Clifford, Can flax replace E-glass in
structural composites? A small wind turbine blade case study,
Composites Part B: Engineering. 52 (2013) 172–181.
doi:10.1016/j.compositesb.2013.04.027.

[47] F. Bensadoun, In-service behaviour of flax fibre reinforced composites


for high performance applications, Ph.D., KU Leuven, 2016.

[48] C. W.Hock, Microscopic structure of flax and related bast fibers,


Journal of Research of the National Bureau of Standard. 29 (1942) 41–
50.

[49] K. Van de Velde, E. Baetens, Thermal and Mechanical Properties of


Flax Fibres as Potential Composite Reinforcement, Macromolecular
Materials and Engineering. 286 (2001) 342–349. doi:10.1002/1439.

[50] H.L. Bos, M.J.A.V.D. Oever, O.C.J.J. Peters, Tensile and compressive
properties of flax fibres for natural fibre reinforced composites, Journal
of Materials Science. 37 (2002) 1683–1692.
doi:10.1023/A:1014925621252.

[51] H.L. Bos, J. Müssig, M.J.A. van den Oever, Mechanical properties of
short-flax-fibre reinforced compounds, Composites Part A: Applied
Science and Manufacturing. 37 (2006) 1591–1604.
doi:10.1016/j.compositesa.2005.10.011.

[52] K. Charlet, C. Baley, C. Morvan, J.P. Jernot, M. Gomina, J. Bréard,


Characteristics of Hermès flax fibres as a function of their location in
the stem and properties of the derived unidirectional composites,
Composites Part A: Applied Science and Manufacturing. 38 (2007)
1912–1921. doi:10.1016/j.compositesa.2007.03.006.

[53] W. Hamad, Cellulosic Materials: Fibers, Networks, and Composites,


Springer Science & Business Media, 2001.

[54] X.S. Sun, Overview of Plant Polymers, in: Handbook of Biopolymers


and Biodegradable Plastics, Elsevier, 2013: pp. 1–10.

[55] O. Faruk, M. Sain, Biofiber Reinforcements in Composite Materials,


Elsevier, 2014.

[56] A. Stamboulis, C.A. Baillie, T. Peijs, Effects of environmental


conditions on mechanical and physical properties of flax fibers,
Composites Part A: Applied Science and Manufacturing. 32 (2001)
1105–1115. doi:10.1016/S1359-835X(01)00032-X.

[57] A.K. Bledzki, J. Gassan, Composites reinforced with cellulose based


fibres, Progress in Polymer Science. 24 (1999) 221–274.
doi:10.1016/S0079-6700(98)00018-5.
204 References

[58] D.N. Saheb, J.P. Jog, Natural fiber polymer composites: a review,
Advances in Polymer Technology. 18 (1999) 351–363.

[59] H.P.S.A. Khalil, H.D. Rozman, M.N. Ahmad, H. Ismail, Acetylated


Plant-Fiber-Reinforced Polyester Composites: A Study of Mechanical,
Hygrothermal, and Aging Characteristics, Polymer-Plastics Technology
and Engineering. 39 (2000) 757–781. doi:10.1081/PPT-100100057.

[60] C. Baley, F. Busnel, Y. Grohens, O. Sire, Influence of chemical


treatments on surface properties and adhesion of flax fibre–polyester
resin, Composites Part A: Applied Science and Manufacturing. 37
(2006) 1626–1637. doi:10.1016/j.compositesa.2005.10.014.

[61] E. Trujillo, M. Moesen, L. Osorio, A.W. Van Vuure, J. Ivens, I.


Verpoest, Bamboo fibres for reinforcement in composite materials:
Strength Weibull analysis, Composites Part A: Applied Science and
Manufacturing. 61 (2014) 115–125.
doi:10.1016/j.compositesa.2014.02.003.

[62] A. Komuraiah, N.S. Kumar, B.D. Prasad, Chemical Composition of


Natural Fibers and its Influence on their Mechanical Properties, Mech
Compos Mater. (2014) 1–18. doi:10.1007/s11029-014-9422-2.

[63] J. Baets, D. Plastria, J. Ivens, I. Verpoest, Determination of the optimal


flax fibre preparation for use in unidirectional flax–epoxy composites,
Journal of Reinforced Plastics and Composites. 33 (2014) 493–502.
doi:10.1177/0731684413518620.

[64] M. Assarar, D. Scida, A. El Mahi, C. Poilâne, R. Ayad, Influence of


water ageing on mechanical properties and damage events of two
reinforced composite materials: Flax–fibres and glass–fibres, Materials
& Design. 32 (2011) 788–795. doi:10.1016/j.matdes.2010.07.024.

[65] A. Bourmaud, C. Baley, Effects of thermo mechanical processing on the


mechanical properties of biocomposite flax fibers evaluated by
nanoindentation, Polymer Degradation and Stability. 95 (2010) 1488–
1494. doi:10.1016/j.polymdegradstab.2010.06.022.

[66] N.P.J. Dissanayake, J. Summerscales, S.M. Grove, M.M. Singh, Life


Cycle Impact Assessment of Flax Fibre for the Reinforcement of
Composites, Journal of Biobased Materials and Bioenergy. 3 (2009)
245–248. doi:10.1166/jbmb.2009.1029.

[67] A. Arbelaiz, B. Fernández, G. Cantero, R. Llano-Ponte, A. Valea, I.


Mondragon, Mechanical properties of flax fibre/polypropylene
composites. Influence of fibre/matrix modification and glass fibre
hybridization, Composites Part A: Applied Science and Manufacturing.
36 (2005) 1637–1644. doi:10.1016/j.compositesa.2005.03.021.
References 205

[68] A.K. Bledzki, The effects of acetylation on properties of flax fibre and
its polypropylene composites, eXPRESS Polymer Letters. 2 (2008)
413–422. doi:10.3144/expresspolymlett.2008.50.

[69] S. Marais, F. Gouanvé, A. Bonnesoeur, J. Grenet, F. Poncin-Epaillard,


C. Morvan, M. Métayer, Unsaturated polyester composites reinforced
with flax fibers: effect of cold plasma and autoclave treatments on
mechanical and permeation properties, Composites Part A: Applied
Science and Manufacturing. 36 (2005) 975–986.
doi:10.1016/j.compositesa.2004.11.008.

[70] A.K. Bledzki, O. Faruk, V.E. Sperber, Cars from Bio-Fibres,


Macromol. Mater. Eng. 291 (2006) 449–457.
doi:10.1002/mame.200600113.

[71] J. Baets, F. Vanfleteren, I. Verpoest, UD-flax preforms for optimal


natural fibre composites performance, Status: Published. (2013) 1–6.

[72] S. Goutianos, T. Peijs, The optimisation of flax fibre yarns for the
development of high-performance natural fibre composites., Advanced
Composites Letters. 12 (2003) 237–241.

[73] G. Huang, L. Liu, Research on properties of thermoplastic composites


reinforced by flax fabrics, Materials & Design. 29 (2008) 1075–1079.
doi:10.1016/j.matdes.2007.03.027.

[74] D.U. Shah, P.J. Schubel, M.J. Clifford, Modelling the effect of yarn
twist on the tensile strength of unidirectional plant fibre yarn
composites, Journal of Composite Materials. 47 (2013) 425–436.

[75] B. Weager, Growing for Gold, Materials World. 21 (2013) 24–25.

[76] S.K. Mazumdar, Composites manufacturing: materials, product, and


process engineering, CRC Press, Boca Raton, Fla., 2002.

[77] M.S. Huda, L.T. Drzal, A.K. Mohanty, M. Misra, Effect of fiber
surface-treatments on the properties of laminated biocomposites from
poly(lactic acid) (PLA) and kenaf fibers, Composites Science and
Technology. 68 (2008) 424–432.
doi:10.1016/j.compscitech.2007.06.022.

[78] O. Faruk, A.K. Bledzki, H.-P. Fink, M. Sain, Biocomposites reinforced


with natural fibers: 2000–2010, Progress in Polymer Science. 37 (2012)
1552–1596. doi:10.1016/j.progpolymsci.2012.04.003.

[79] S. Huo, A. Thapa, C.A. Ulven, Effect of surface treatments on


interfacial properties of flax fiber-reinforced composites, Advanced
Composite Materials. 22 (2013) 109–121.
doi:10.1080/09243046.2013.777996.
206 References

[80] G. Francucci, E. Rodriguez, Processing of plant fiber composites by


liquid molding techniques: An overview, Polym. Compos. (2014) n/a-
n/a. doi:10.1002/pc.23229.

[81] D30 Committee, Test Method for Measuring the Damage Resistance of
a Fiber-Reinforced Polymer Matrix Composite to a Drop-Weight
Impact Event, ASTM International, 2012.

[82] K.T. Tan, N. Watanabe, Y. Iwahori, T. Ishikawa, Effect of stitch density


and stitch thread thickness on compression after impact strength and
response of stitched composites, Composites Science and Technology.
72 (2012) 587–598. doi:10.1016/j.compscitech.2012.01.003.

[83] I.K. Giannopoulos, E.E. Theotokoglou, X. Zhang, Impact damage and


CAI strength of a woven CFRP material with fire retardant properties,
Composites Part B: Engineering. 91 (2016) 8–17.
doi:10.1016/j.compositesb.2016.01.041.

[84] W. Tan, B.G. Falzon, L.N.S. Chiu, M. Price, Predicting low velocity
impact damage and Compression-After-Impact (CAI) behaviour of
composite laminates, Composites Part A: Applied Science and
Manufacturing. 71 (2015) 212–226.
doi:10.1016/j.compositesa.2015.01.025.

[85] G.A. Bibo, P.J. Hogg, Influence of reinforcement architecture on


damage mechanisms and residual strength of glass-fibre/epoxy
composite systems, Composites Science and Technology. 58 (1998)
803–813. doi:10.1016/S0266-3538(97)00055-9.

[86] G. Caprino, V. Lopresto, On the penetration energy for fibre-reinforced


plastics under low-velocity impact conditions, Composites Science and
Technology. 61 (2001) 65–73. doi:10.1016/S0266-3538(00)00152-4.

[87] T.-W. Shyr, Y.-H. Pan, Impact resistance and damage characteristics of
composite laminates, Composite Structures. 62 (2003) 193–203.
doi:10.1016/S0263-8223(03)00114-4.

[88] M. Aktaş, H. Ersen Balcıoğlu, A. Aktaş, E. Türker, M. Emin Deniz,


Impact and post impact behavior of layer fabric composites, Composite
Structures. 94 (2012) 2809–2818.
doi:10.1016/j.compstruct.2012.04.008.

[89] B. Schrauwen, T. Peijs, Influence of Matrix Ductility and Fibre


Architecture on the Repeated Impact Response of Glass-Fibre-
Reinforced Laminated Composites, Applied Composite Materials. 9
(2002) 331–352. doi:10.1023/A:1020267013414.

[90] N.F. Rilo, L.M.S. Ferreira, Experimental study of low-velocity impacts


on glass-epoxy laminated composite plates, Int J Mech Mater Des. 4
(2008) 291–300. doi:10.1007/s10999-008-9071-5.
References 207

[91] P.O. Sjoblom, J.T. Hartness, T.M. Cordell, On Low-Velocity Impact


Testing of Composite Materials, Journal of Composite Materials. 22
(1988) 30–52. doi:10.1177/002199838802200103.

[92] ISO 6603-2:2000 - Plastics: Determination of puncture impact


behaviour of rigid plastics, Part 2: Instrumented impact testing, BSI,
2002.

[93] H.N. Dhakal, Z.Y. Zhang, M.O.W. Richardson, O.A.Z. Errajhi, The low
velocity impact response of non-woven hemp fibre reinforced
unsaturated polyester composites, Composite Structures. 81 (2007)
559–567.

[94] C. Santulli, A.P. Caruso, Effect of Fibre Architecture on the Falling


Weight Impact Properties of Hemp/Epoxy Composites, Journal of
Biobased Materials and Bioenergy. 3 (2009) 291–297.

[95] C. Scarponi, C.S. Pizzinelli, S. Sánchez-Sáez, E. Barbero, Impact load


behaviour of resin transfer moulding (RTM) hemp fibre composite
laminates, Journal of Biobased Materials and Bioenergy. 3 (2009) 298–
310.

[96] H.N. Dhakal, Z.Y. Zhang, N. Bennett, P.N.B. Reis, Low-velocity


impact response of non-woven hemp fibre reinforced unsaturated
polyester composites: Influence of impactor geometry and impact
velocity, Composite Structures. 94 (2012) 2756–2763.

[97] D.S. de Vasconcellos, F. Sarasini, F. Touchard, L. Chocinski-Arnault,


M. Pucci, C. Santulli, J. Tirillò, S. Iannace, L. Sorrentino, Influence of
low velocity impact on fatigue behaviour of woven hemp fibre
reinforced epoxy composites, Composites Part B: Engineering. 66
(2014) 46–57.

[98] H.N. Dhakal, V. Arumugam, A. Aswinraj, C. Santulli, Z.Y. Zhang, A.


Lopez-Arraiza, Influence of temperature and impact velocity on the
impact response of jute/UP composites, Polymer Testing. 35 (2014) 10–
19.

[99] H.N. Dhakal, M. Skrifvars, K. Adekunle, Z.Y. Zhang, Falling weight


impact response of jute/methacrylated soybean oil bio-composites under
low velocity impact loading, Composites Science and Technology. 92
(2014) 134–141.

[100] G. Caprino, L. Carrino, M. Durante, A. Langella, V. Lopresto, Low


impact behaviour of hemp fibre reinforced epoxy composites,
Composite Structures. 133 (2015) 892–901.

[101] P. Wambua, B. Vangrimde, S. Lomov, I. Verpoest, The response of


natural fibre composites to ballistic impact by fragment simulating
projectiles, Composite Structures. 77 (2007) 232–240.
208 References

[102] H.T.N. Kuan, W.J. Cantwell, M.A. Hazizan, C. Santulli, The fracture
properties of environmental-friendly fiber metal laminates, Journal of
Reinforced Plastics and Composites. 30 (2011) 499–508.
doi:10.1177/0731684411398536.

[103] R. Velmurugan, S. Solaimurugan, Improvements in Mode I interlaminar


fracture toughness and in-plane mechanical properties of stitched
glass/polyester composites, Composites Science and Technology. 67
(2007) 61–69. doi:10.1016/j.compscitech.2006.03.032.

[104] Muruganandhan, V. Murali, Mode-I Fracture and Impact Analysis on


Stitched and Unstitched Glass/Epoxy Composite Laminate, Procedia
Engineering. 38 (2012) 2207–2213.

[105] T.J. Kang, S.H. Lee, Effect of Stitching on the Mechanical and Impact
Properties of Woven Laminate Composite, Journal of Composite
Materials. 28 (1994) 1574–1587.

[106] A.P. Mouritz, K.H. Leong, I. Herszberg, A review of the effect of


stitching on the in-plane mechanical properties of fibre-reinforced
polymer composites, Composites Part A: Applied Science and
Manufacturing. 28 (1997) 979–991.

[107] A. Miravete, 3-D textile reinforcements in composite materials,


Woodhead Publishing Ltd, Cambridge, 1999.

[108] S. Solaimurugan, R. Velmurugan, Influence of in-plane fibre orientation


on mode I interlaminar fracture toughness of stitched glass/polyester
composites, Composites Science and Technology. 68 (2008) 1742–
1752. doi:10.1016/j.compscitech.2008.02.008.

[109] K. Dransfield, C. Baillie, Y.-W. Mai, Improving the delamination


resistance of CFRP by stitching—a review, Composites Science and
Technology. 50 (1994) 305–317. doi:10.1016/0266-3538(94)90019-1.

[110] L.K. Jain, Y.-W. Mai, On the effect of stitching on mode I delamination
toughness of laminated composites, Composites Science and
Technology. 51 (1994) 331–345. doi:10.1016/0266-3538(94)90103-1.

[111] A. Weinberg, Y. Eilam, Needle for machine stitching with a composite


thread, US7487733 B2, 2009.

[112] A. Weinberg, Y. Eilam, Sewing machine for stitching with a composite


thread, US7475647 B2, 2009.

[113] T. Tan, C. Sun, L. Mignery, The use of stitching to suppress


delamination in laminated composites, Delamination and Debonding of
materials(A 86-20626 07-24). Philadelphia, PA, American Society for
Testing and Materials, 1985,. (1985).
References 209

[114] K.P. Plain, L. Tong, An experimental study on mode I and II fracture


toughness of laminates stitched with a one-sided stitching technique,
Composites Part A: Applied Science and Manufacturing. 42 (2011)
203–210. doi:10.1016/j.compositesa.2010.11.006.

[115] K.T. Tan, N. Watanabe, Y. Iwahori, Stitch fiber comparison for


improvement of interlaminar fracture toughness in stitched composites,
Journal of Reinforced Plastics and Composites. 30 (2011) 99–109.
doi:10.1177/0731684410383065.

[116] W.C. Chung, B.Z. Jang, T.C. Chang, L.R. Hwang, R.C. Wilcox,
Fracture behavior in stitched multidirectional composites, Materials
Science and Engineering: A. 112 (1989) 157–173. doi:10.1016/0921-
5093(89)90355-9.

[117] M.D.K. Wood, X. Sun, L. Tong, A. Katzos, A.R. Rispler, Y.-W. Mai,
The effect of stitch distribution on Mode I delamination toughness of
stitched laminated composites – experimental results and FEA
simulation, Composites Science and Technology. 67 (2007) 1058–1072.
doi:10.1016/j.compscitech.2006.06.002.

[118] X. Sun, M.D.K. Wood, L. Tong, A parametric study on the design of


stitched laminated DCB specimens, Composite Structures. 75 (2006)
72–78. doi:10.1016/j.compstruct.2006.04.061.

[119] N.E. Vandermey, D.H. Morris, J.E. Masters, Damage development


under compression-compression fatigue loading in a stitched uniwoven
graphite/epoxy composite material, (1991).

[120] E. Wu, J. Wang, Behavior of Stitched Laminates under In-Plane Tensile


and Transverse Impact Loading, Journal of Composite Materials. 29
(1995) 2254–2279. doi:10.1177/002199839502901702.

[121] L.K. Jain, K.A. Dransfield, Y.-W. Mai, On the effects of stitching in
CFRPs—II. Mode II delamination toughness, Composites Science and
Technology. 58 (1998) 829–837.

[122] K.T. Tan, N. Watanabe, Y. Iwahori, Experimental investigation of


bridging law for single stitch fibre using Interlaminar tension test,
Composite Structures. 92 (2010) 1399–1409.
doi:10.1016/j.compstruct.2009.11.018.

[123] H. Heß, N. Himmel, Structurally stitched NCF CFRP laminates. Part 1:


Experimental characterization of in-plane and out-of-plane properties,
Composites Science and Technology. 71 (2011) 549–568.

[124] F. Aymerich, R. Onnis, P. Priolo, Analysis of the effect of stitching on


the fatigue strength of single-lap composite joints, Composites Science
and Technology. 66 (2006) 166–175.
doi:10.1016/j.compscitech.2005.04.023.
210 References

[125] K. Bilisik, Three-dimensional braiding for composites: A review,


Textile Research Journal. 83 (2013) 1414–1436.
doi:10.1177/0040517512450766.

[126] A.P. Mouritz, B.N. Cox, A mechanistic interpretation of the


comparative in-plane mechanical properties of 3D woven, stitched and
pinned composites, Composites Part A: Applied Science and
Manufacturing. 41 (2010) 709–728.
doi:10.1016/j.compositesa.2010.02.001.

[127] A.P. Mouritz, Tensile fatigue properties of 3D composites with through-


thickness reinforcement, Composites Science and Technology. 68
(2008) 2503–2510. doi:10.1016/j.compscitech.2008.05.003.

[128] V. Lopresto, V. Melito, C. Leone, G. Caprino, Effect of stitches on the


impact behaviour of graphite/epoxy composites, Composites Science
and Technology. 66 (2006) 206–214.

[129] K.T. Tan, N. Watanabe, Y. Iwahori, Effect of stitch density and stitch
thread thickness on low-velocity impact damage of stitched composites,
Composites Part A: Applied Science and Manufacturing. 41 (2010)
1857–1868.

[130] X. Guo, H. Long, L. Zhao, Investigation on the impact and


compression-after-impact properties of warp-knitted spacer fabrics,
Textile Research Journal. 83 (2013) 904–916.
doi:10.1177/0040517512461682.

[131] A. Yudhanto, N. Watanabe, Y. Iwahori, H. Hoshi, Effect of stitch


density on tensile properties and damage mechanisms of stitched
carbon/epoxy composites, Composites Part B: Engineering. 46 (2013)
151–165.

[132] K.T. Tan, N. Watanabe, Y. Iwahori, X-ray radiography and micro-


computed tomography examination of damage characteristics in
stitched composites subjected to impact loading, Composites Part B:
Engineering. 42 (2011) 874–884.
doi:10.1016/j.compositesb.2011.01.011.

[133] A. Yoshimura, T. Nakao, S. Yashiro, N. Takeda, Improvement on out-


of-plane impact resistance of CFRP laminates due to through-the-
thickness stitching, Composites Part A: Applied Science and
Manufacturing. 39 (2008) 1370–1379.
doi:10.1016/j.compositesa.2008.04.019.

[134] J.-H. Byun, S.-W. Song, C.-H. Lee, M.-K. Um, B.-S. Hwang, Impact
properties of laminated composites with stitching fibers, Composite
Structures. 76 (2006) 21–27. doi:10.1016/j.compstruct.2006.06.004.
References 211

[135] A.P. Mouritz, J. Gallagher, A.A. Goodwin, Flexural strength and


interlaminar shear strength of stitched GRP laminates following
repeated impacts, Composites Science and Technology. 57 (1997) 509–
522. doi:10.1016/S0266-3538(96)00164-9.

[136] N. Ghafari-Namini, H. Ghasemnejad, Effect of natural stitched


composites on the crashworthiness of box structures, Materials &
Design. 39 (2012) 484–494. doi:10.1016/j.matdes.2012.03.025.

[137] H.N. Dhakal, Z.Y. Zhang, R. Guthrie, J. MacMullen, N. Bennett,


Development of flax/carbon fibre hybrid composites for enhanced
properties, Carbohydrate Polymers. 96 (2013) 1–8.
doi:10.1016/j.carbpol.2013.03.074.

[138] J. Flynn, A. Amiri, C. Ulven, Hybridized carbon and flax fiber


composites for tailored performance, Materials & Design. 102 (2016)
21–29. doi:10.1016/j.matdes.2016.03.164.

[139] J. Modniks, J. Andersons, Modeling the non-linear deformation of a


short-flax-fiber-reinforced polymer composite by orientation averaging,
Composites Part B: Engineering. 54 (2013) 188–193.
doi:10.1016/j.compositesb.2013.04.058.

[140] J. Sliseris, L. Yan, B. Kasal, Numerical modelling of flax short fibre


reinforced and flax fibre fabric reinforced polymer composites,
Composites Part B: Engineering. 89 (2016) 143–154.
doi:10.1016/j.compositesb.2015.11.038.

[141] C. Mattrand, A. Béakou, K. Charlet, Numerical modeling of the flax


fiber morphology variability, Composites Part A: Applied Science and
Manufacturing. 63 (2014) 10–20.
doi:10.1016/j.compositesa.2014.03.020.

[142] A. Beakou, K. Charlet, Mechanical properties of interfaces within a flax


bundle—Part II: Numerical analysis, International Journal of Adhesion
and Adhesives. 43 (2013) 54–59. doi:10.1016/j.ijadhadh.2013.01.013.

[143] C. Poilâne, Z.E. Cherif, F. Richard, A. Vivet, B. Ben Doudou, J. Chen,


Polymer reinforced by flax fibres as a viscoelastoplastic material,
Composite Structures. 112 (2014) 100–112.
doi:10.1016/j.compstruct.2014.01.043.

[144] A. Rubio-López, T. Hoang, C. Santiuste, Constitutive model to predict


the viscoplastic behaviour of natural fibres based composites,
Composite Structures. 155 (2016) 8–18.
doi:10.1016/j.compstruct.2016.08.001.

[145] D. Shu, Y.-W. Mai, Effect of stitching on interlaminar delamination


extension in composite laminates, Composites Science and Technology.
49 (1993) 165–171.
212 References

[146] R. Massabo, B.N. Cox, Concepts for bridged Mode ii delamination


cracks, Journal of the Mechanics and Physics of Solids. 47 (1999)
1265–1300. doi:10.1016/S0022-5096(98)00107-0.

[147] H. Heß, N. Himmel, Structurally stitched NCF CFRP laminates. Part 2:


Finite element unit cell based prediction of in-plane strength,
Composites Science and Technology. 71 (2011) 569–585.

[148] B.V. Sankar, S.M. Dharmapuri, Analysis of a Stitched Double


Cantilever Beam, Journal of Composite Materials. 32 (1998) 2203–
2225. doi:10.1177/002199839803202402.

[149] L. Chen, P.G. Ifju, B.V. Sankar, Analysis of Mode I and Mode II Tests
for Composites with Translaminar Reinforcements, Journal of
Composite Materials. 39 (2005) 1311–1333.
doi:10.1177/0021998305050425.

[150] B.F. Sørensen, T.K. Jacobsen, Large-scale bridging in composites: R-


curves and bridging laws, Composites Part A: Applied Science and
Manufacturing. 29 (1998) 1443–1451. doi:10.1016/S1359-
835X(98)00025-6.

[151] A. Airoldi, C.G. Dávila, Identification of material parameters for


modelling delamination in the presence of fibre bridging, Composite
Structures. 94 (2012) 3240–3249.
doi:10.1016/j.compstruct.2012.05.014.

[152] V. Tamuzs, S. Tarasovs, U. Vilks, Progressive delamination and fiber


bridging modeling in double cantilever beam composite specimens,
Engineering Fracture Mechanics. 68 (2001) 513–525.
doi:10.1016/S0013-7944(00)00131-4.

[153] L. Sorensen, J. Botsis, T. Gmür, L. Humbert, Bridging tractions in


mode I delamination: Measurements and simulations, Composites
Science and Technology. 68 (2008) 2350–2358.
doi:10.1016/j.compscitech.2007.08.024.

[154] C.G. Dávila, C.A. Rose, P.P. Camanho, A procedure for superposing
linear cohesive laws to represent multiple damage mechanisms in the
fracture of composites, Int J Fract. 158 (2009) 211–223.
doi:10.1007/s10704-009-9366-z.

[155] B.F. Sørensen, S. Goutianos, T.K. Jacobsen, Strength scaling of


adhesive joints in polymer–matrix composites, International Journal of
Solids and Structures. 46 (2009) 741–761.
doi:10.1016/j.ijsolstr.2008.09.024.

[156] D. Scida, Z. Aboura, M.L. Benzeggagh, E. Bocherens, A


micromechanics model for 3D elasticity and failure of woven-fibre
References 213

composite materials, Composites Science and Technology. 59 (1999)


505–517. doi:10.1016/S0266-3538(98)00096-7.

[157] O. Cousigné, D. Moncayo, D. Coutellier, P. Camanho, H. Naceur,


Numerical modeling of nonlinearity, plasticity and damage in CFRP-
woven composites for crash simulations, Composite Structures. 115
(2014) 75–88. doi:10.1016/j.compstruct.2014.04.017.

[158] A.F. Johnson, Modelling fabric reinforced composites under impact


loads, Composites Part A: Applied Science and Manufacturing. 32
(2001) 1197–1206. doi:10.1016/S1359-835X(00)00186-X.

[159] O. Cousigné, D. Moncayo, D. Coutellier, P. Camanho, H. Naceur, S.


Hampel, Development of a new nonlinear numerical material model for
woven composite materials accounting for permanent deformation and
damage, Composite Structures. 106 (2013) 601–614.
doi:10.1016/j.compstruct.2013.07.026.

[160] L. Iannucci, Progressive failure modelling of woven carbon composite


under impact, International Journal of Impact Engineering. 32 (2006)
1013–1043. doi:10.1016/j.ijimpeng.2004.08.006.

[161] P.S. Shembekar, N.K. Naik, Elastic Behavior of Woven Fabric


Composites: II—Laminate Analysis, Journal of Composite Materials.
26 (1992) 2226–2246. doi:10.1177/002199839202601503.

[162] X. Tang, J.D. Whitcomb, Progressive Failure Behaviors of 2D Woven


Composites, Journal of Composite Materials. 37 (2003) 1239–1259.
doi:10.1177/0021998303037014002.

[163] M. Zako, Y. Uetsuji, T. Kurashiki, Finite element analysis of damaged


woven fabric composite materials, Composites Science and
Technology. 63 (2003) 507–516. doi:10.1016/S0266-3538(02)00211-7.

[164] J. LLorca, C. González, J.M. Molina-Aldareguía, J. Segurado, R.


Seltzer, F. Sket, M. Rodríguez, S. Sádaba, R. Muñoz, L.P. Canal,
Multiscale Modeling of Composite Materials: a Roadmap Towards
Virtual Testing, Adv. Mater. 23 (2011) 5130–5147.
doi:10.1002/adma.201101683.

[165] J.Z. Mao, X.S. Sun, M. Ridha, V.B.C. Tan, T.E. Tay, A Modeling
Approach Across Length Scales for Progressive Failure Analysis of
Woven Composites, Applied Composite Materials. 20 (2012) 213–231.
doi:10.1007/s10443-012-9266-7.

[166] S.V. Lomov, D.S. Ivanov, I. Verpoest, M. Zako, T. Kurashiki, H.


Nakai, S. Hirosawa, Meso-FE modelling of textile composites: Road
map, data flow and algorithms, Composites Science and Technology. 67
(2007) 1870–1891. doi:10.1016/j.compscitech.2006.10.017.
214 References

[167] G. Ernst, M. Vogler, C. Hühne, R. Rolfes, Multiscale progressive


failure analysis of textile composites, Composites Science and
Technology. 70 (2010) 61–72. doi:10.1016/j.compscitech.2009.09.006.

[168] Composites Evolution, (n.d.). http://compositesevolution.com/


(accessed October 15, 2016).

[169] LINEO - Advanced Flax, (n.d.). http://www.lineo.eu/ (accessed October


15, 2016).

[170] Cotton Threads, (n.d.).


https://www.guetermann.com/shop/en/view/content/Products-
Cotton?node=Consumer-Cotton-Threads (accessed October 15, 2016).

[171] Axson | Axson, (n.d.). http://www.axson-


technologies.com/us/index.html (accessed October 15, 2016).

[172] D20 Committee, Test Methods for Flexural Properties of Unreinforced


and Reinforced Plastics and Electrical Insulating Materials, ASTM
International, 2010.

[173] D30 Committee, Test Method for Mode I Interlaminar Fracture


Toughness of Unidirectional Fiber-Reinforced Polymer Matrix
Composites, ASTM International, 2007.

[174] Y. Ogo, The effect of stitching on in-plane and interlaminar properties


of carbon-epoxy fabric laminates, CCM Report Number 87-17, Center
for Composite Materials, University of Delaware, Newark, pp 1-188,
1987.

[175] M. Olave, I. Vara, H. Husabiaga, L. Aretxabaleta, S.V. Lomov, D.


Vandepitte, Nesting effect on the mode I fracture toughness of woven
laminates, Composites Part A: Applied Science and Manufacturing. 74
(2015) 166–173. doi:10.1016/j.compositesa.2015.03.017.

[176] R.A. Pearson, A.F. Yee, Toughening mechanisms in thermoplastic-


modified epoxies: 1. Modification using poly(phenylene oxide),
Polymer. 34 (1993) 3658–3670.

[177] T. Ebeling, A. Hiltner, E. Baer, I.M. Fraser, M.L. Orton, Delamination


Failure of a Woven Glass Fiber Composite, Journal of Composite
Materials. 31 (1997) 1318–1333.

[178] J.-K. Kim, M.-L. Sham, Impact and delamination failure of woven-
fabric composites, Composites Science and Technology. 60 (2000)
745–761.

[179] C. Boyang, Numerical Modelling of Scale-Dependent Damage and


Failure of Composites, Ph.D., 2013.
References 215

[180] M. Ravandi, W.S. Teo, L.Q.N. Tran, M.S. Yong, T.E. Tay, The effects
of through-the-thickness stitching on the Mode I interlaminar fracture
toughness of flax/epoxy composite laminates, Materials & Design. 109
(2016) 659–669. doi:10.1016/j.matdes.2016.07.093.

[181] A. Turon, C.G. Dávila, P.P. Camanho, J. Costa, An engineering solution


for mesh size effects in the simulation of delamination using cohesive
zone models, Engineering Fracture Mechanics. 74 (2007) 1665–1682.
doi:10.1016/j.engfracmech.2006.08.025.

[182] B.Y. Chen, T.E. Tay, P.M. Baiz, S.T. Pinho, Numerical analysis of size
effects on open-hole tensile composite laminates, Composites Part A:
Applied Science and Manufacturing. 47 (2013) 52–62.
doi:10.1016/j.compositesa.2012.12.001.

[183] L. Lampani, Finite element analysis of delamination of a composite


component with the cohesive zone model technique, Engineering
Computations. 28 (2011) 30–46.
doi:http://dx.doi.org.libproxy1.nus.edu.sg/10.1108/02644401111097000
.

[184] P.P. Camanho, C.G. Dávila, Mixed-mode decohesion finite elements for
the simulation of delamination in composite materials, NASA-
Technical Paper. 211737 (2002) 33.

[185] Abaqus Analysis User’s Manual, Version 6.11, 2011.

[186] M.L. Benzeggagh, M. Kenane, Measurement of mixed-mode


delamination fracture toughness of unidirectional glass/epoxy
composites with mixed-mode bending apparatus, Composites Science
and Technology. 56 (1996) 439–449. doi:10.1016/0266-
3538(96)00005-X.

[187] M. Ravandi, W.S. Teo, L.Q.N. Tran, M.S.Y. and T.E. Tay, Mode I
Interlaminar Fracture Toughness of Natural Fiber Stitched Flax/Epoxy
Composite Laminates – Experimental and Numerical Analysis,
Proceedings of the American Society for Composites: Thirty-First
Technical Conference. 0 (2016). http://dpi-
proceedings.com/index.php/asc31/article/view/3287 (accessed October
4, 2016).

[188] C.H. Popelar, M.F. Kanninen, Advanced Fracture Mechanics, 1 edition,


Oxford University Press, New York, 1985.

[189] I.N. Dyson, A.J. Kinloch, A. Okada, The interlaminar failure behaviour
of carbon fibre/polyetheretherketone composites, Composites. 25
(1994) 189–196.

[190] M. Ravandi, W.S. Teo, L.Q.N. Tran, M.S. Yong, T.E. Tay, Effect of
through-the-thickness stitching and fibre architecture on the
216 References

interlaminar properties of flax/epoxy laminates, in: Proceedings of the


20th International Conference on Composite Materials ICCM-20,
Copenhagen, Denmark, July 19-24, 2015.

[191] P.P. Camanho, C.G. Davila, M.F. De Moura, Numerical simulation of


mixed-mode progressive delamination in composite materials, Journal
of Composite Materials. 37 (2003) 1415–1438.

[192] M. Ravandi, W.S. Teo, L.Q.N. Tran, M.S. Yong, T.E. Tay, Low
velocity impact performance of stitched flax/epoxy composite
laminates, Composites Part B: Engineering. 117 (2017) 89–100.
doi:10.1016/j.compositesb.2017.02.003.

[193] M. Ravandi, W.S. Teo, L.Q.N. Tran, M.S. Yong, T.E. Tay, Effects of
the natural fiber stitching on the impact and compression after impact
response of the flax/epoxy composites, in: Munich, Germany, 2016.

[194] F. Aymerich, F. Dore, P. Priolo, Prediction of impact-induced


delamination in cross-ply composite laminates using cohesive interface
elements, Composites Science and Technology. 68 (2008) 2383–2390.
doi:10.1016/j.compscitech.2007.06.015.

[195] Y. Xue, Y. Du, S. Elder, K. Wang, J. Zhang, Temperature and loading


rate effects on tensile properties of kenaf bast fiber bundles and
composites, Composites Part B: Engineering. 40 (2009) 189–196.
doi:10.1016/j.compositesb.2008.11.009.

[196] S. Abrate, Impact on Laminated Composite Materials, Appl. Mech.


Rev. 44 (1991) 155–190. doi:10.1115/1.3119500.

[197] http://www.axson-na.com/TDSs/TDS - Epolam 5015 System, Epolam


5015-5014-5016, (n.d.).

[198] P.W. Harper, S.R. Hallett, Cohesive zone length in numerical


simulations of composite delamination, Engineering Fracture
Mechanics. 75 (2008) 4774–4792.
doi:10.1016/j.engfracmech.2008.06.004.

[199] F. Pascal, P. Navarro, S. Marguet, J.-F. Ferrero, On the modelling of


low to medium velocity impact onto woven composite materials with a
2D semi-continuous approach, Composite Structures. 134 (2015) 302–
310. doi:10.1016/j.compstruct.2015.08.067.

[200] P. Sztefek, R. Olsson, Nonlinear compressive stiffness in impacted


composite laminates determined by an inverse method, Composites Part
A: Applied Science and Manufacturing. 40 (2009) 260–272.
doi:10.1016/j.compositesa.2008.12.002.
References 217

[201] V.A. Guénon, T.-W. Chou, J.W.G. Jr, Toughness properties of a three-
dimensional carbon-epoxy composite, J Mater Sci. 24 (1989) 4168–
4175. doi:10.1007/BF01168991.

[202] J.-H. Byun, J.W. Gillespie, T.-W. Chou, Mode I Delamination of a


Three-Dimensional Fabric Composite, Journal of Composite Materials.
24 (1990) 497–518. doi:10.1177/002199839002400503.

[203] J.M. Rice, Y.K. Kim, A.F. Lewis, B.K. Fink, Fracture Toughness of
Through-Thickness Reinforced Composites, NTC Project: F04-MD12,
National Textile Center Annual Report. (2004) 1–7.

[204] I. Van de Weyenberg, T. Chi Truong, B. Vangrimde, I. Verpoest,


Improving the properties of UD flax fibre reinforced composites by
applying an alkaline fibre treatment, Composites Part A: Applied
Science and Manufacturing. 37 (2006) 1368–1376.
doi:10.1016/j.compositesa.2005.08.016.

[205] J. George, J.I. Verpoest I., Mechanical properties of flax fibre


reinforced epoxy composites, Die Angewandte Makromolekulare
Chemie. 272 (1999) 41–45. doi:10.1002/(SICI)1522-
9505(19991201)272:1<41::AID-APMC41>3.0.CO;2-X.

[206] I. Van de Weyenberg, J. Ivens, A. De Coster, B. Kino, E. Baetens, I.


Verpoest, Influence of processing and chemical treatment of flax fibres
on their composites, Composites Science and Technology. 63 (2003)
1241–1246. doi:10.1016/S0266-3538(03)00093-9.

[207] D.B. Dittenber, H.V.S. GangaRao, Critical review of recent


publications on use of natural composites in infrastructure, Composites
Part A: Applied Science and Manufacturing. 43 (2012) 1419–1429.
doi:10.1016/j.compositesa.2011.11.019.

[208] G. Kretsis, A review of the tensile, compressive, flexural and shear


properties of hybrid fibre-reinforced plastics, Composites. 18 (1987)
13–23. doi:10.1016/0010-4361(87)90003-6.

[209] I.S. Aji, E.S. Zainudin, K. Abdan, S.M. Sapuan, M.D. Khairul,
Mechanical properties and water absorption behavior of hybridized
kenaf/pineapple leaf fibre-reinforced high-density polyethylene
composite, Journal of Composite Materials. 47 (2013) 979–990.
doi:10.1177/0021998312444147.

[210] F. Sarasini, J. Tirillò, M. Valente, T. Valente, S. Cioffi, S. Iannace, L.


Sorrentino, Effect of basalt fiber hybridization on the impact behavior
under low impact velocity of glass/basalt woven fabric/epoxy resin
composites, Composites Part A: Applied Science and Manufacturing.
47 (2013) 109–123. doi:10.1016/j.compositesa.2012.11.021.
218 References

[211] I.M. De Rosa, Post-impact mechanical characterisation of E-glass/basalt


woven fabric interply hybrid laminates, Express Polymer Letters. 5
(2011) 449–459. doi:10.3144/expresspolymlett.2011.43.

[212] C. Santulli, Impact properties of glass/plant fibre hybrid laminates, J


Mater Sci. 42 (2007) 3699–3707. doi:10.1007/s10853-006-0662-y.

[213] A. Arbelaiz, B. Fernández, G. Cantero, R. Llano-Ponte, A. Valea, I.


Mondragon, Mechanical properties of flax fibre/polypropylene
composites. Influence of fibre/matrix modification and glass fibre
hybridization, Composites Part A: Applied Science and Manufacturing.
36 (2005) 1637–1644. doi:10.1016/j.compositesa.2005.03.021.

[214] M.S. Sreekala, J. George, M.G. Kumaran, S. Thomas, The mechanical


performance of hybrid phenol-formaldehyde-based composites
reinforced with glass and oil palm fibres, Composites Science and
Technology. 62 (2002) 339–353. doi:10.1016/S0266-3538(01)00219-6.

[215] J. Flynn, A. Amiri, C. Ulven, Hybridized carbon and flax fiber


composites for tailored performance, Materials & Design. 102 (2016)
21–29. doi:10.1016/j.matdes.2016.03.164.

[216] H.N. Dhakal, Z.Y. Zhang, R. Guthrie, J. MacMullen, N. Bennett,


Development of flax/carbon fibre hybrid composites for enhanced
properties, Carbohydrate Polymers. 96 (2013) 1–8.
doi:10.1016/j.carbpol.2013.03.074.

[217] Z.S. Bagheri, I. El Sawi, E.H. Schemitsch, R. Zdero, H. Bougherara,


Biomechanical properties of an advanced new carbon/flax/epoxy
composite material for bone plate applications, Journal of the
Mechanical Behavior of Biomedical Materials. 20 (2013) 398–406.
doi:10.1016/j.jmbbm.2012.12.013.

[218] E. Nisini, C. Santulli, A. Liverani, Mechanical and impact


characterization of hybrid composite laminates with carbon, basalt and
flax fibres, Composites Part B: Engineering. (2016).
doi:10.1016/j.compositesb.2016.06.071.

[219] V.A. Patel, B.D. Bhuva, P.H. Parsania, Performance Evaluation of


Treated—Untreated Jute—Carbon and Glass—Carbon Hybrid
Composites of Bisphenol-C based Mixed Epoxy—Phenolic Resins,
Journal of Reinforced Plastics and Composites. 28 (2009) 2549–2556.
doi:10.1177/0731684408093973.

[220] Y. Swolfs, L. Gorbatikh, I. Verpoest, Fibre hybridisation in polymer


composites: A review, Composites Part A: Applied Science and
Manufacturing. 67 (2014) 181–200.
doi:10.1016/j.compositesa.2014.08.027.
References 219

[221] N. Ghafari-Namini, H. Ghasemnejad, Effect of natural stitched


composites on the crashworthiness of box structures, Materials &
Design. 39 (2012) 484–494. doi:10.1016/j.matdes.2012.03.025.

[222] B. Blackman, A. Kinloch, Fracture tests on structural adhesive joints,


in: A.P. and J.G.W. D.R. Moore (Ed.), European Structural Integrity
Society, Elsevier, 2001: pp. 225–267.

[223] D30 Committee, Standard Test Method for In-Plane Shear Response of
Polymer Matrix Composite Materials by Tensile Test of a ±45°
Laminate, ASTM International, 2013.
220 Calculation of Mode I interlaminar fracture toughness (G IC) using DCB test data

Appendix A
Calculation of Mode I interlaminar fracture

toughness (GIC) using DCB test data

This Annex presents three data reduction methods used for calculating

interlaminar fracture toughness (GIC) of un/stitched flax fiber composites

laminates according to ASTM D5528-01 [173]. These methods are known as

modified beam theory (MBT), a compliance calibration method (CC) and a

modified compliance calibration method (MCC). None of these three methods

are clearly superior to the others since the discrepancy in the determined values

of GIC is less than 3.1%. Nevertheless, the MBT method normally yielded the

most conservative values of GIC for 80 % of the specimens tested, as stated by

ASTM D5528.

The data required for the analysis are the crack (delamination) length, a, the

corresponding loads, P, and crack opening displacements, d, which are obtained

from the DCB test. An Initial loading-unloading cycle should also perform to

let the crack propagate 3 to 5 mm from the pre-delamination insert so as to create

a naturally sharp crack tip. The following values should be determined from the

load-displacement curve.

Initiation values:
Appendix A 221

The crack length and the corresponding load and crack opening displacement

of the following initiation values should be determined in order to calculate the

initiation GIC:

o Deviation from linearity (NL): The point of deviation from linearity, is

determined by drawing a straight line from the origin but ignoring any

initial deviations due to the take-up of play in the loading system.

o Visual observation (VIS): The point corresponds to the onset of crack

growth. In other words, the first point at which the crack is observed to

start moving.

o 5% or MAX: The 5% value corresponds to the point on the load-

displacement curve at which the compliance has increased by 5% of its

initial value. A best straight line is drawn to determine the initial

compliance (ignoring any initial deviation). A new line is then drawn

with a compliance equal to additional 5%, and the intersection of this

new line with the load-displacement curve marked. Whichever point

provides maximum value of load that is the point to be used.

Propagation values:

During delamination propagation, the crack length and the corresponding loads

and opening displacements values (PROP in Figure A.1) should also be

determined. These values are used for calculation of propagation GIC.

These imitation and propagation points are illustrated on the load-displacement

curve of Figure A.1. In the case of unstable stick-slip propagation, only the first

initiation value of the crack length and the crack lengths at subsequent arrest
222 Calculation of Mode I interlaminar fracture toughness (G IC) using DCB test data

points can be recorded. For the critical value of the Mode I strain energy release

rate (GIC), the initiation values are taken into consideration (The points ‘i’ in

Figure).

(a) (b)

Figure A.1. (a) Imitation and propagation values on load-displacement curve of the DCB test;
(b) Schematic force-displacement curve for a DCB specimen, exhibiting unstable ‘stick-slip’
crack growth behavior. (‘i’ indicates initiation points, ‘a’ indicates arrest points.) [222].

A.1 Modified Beam Theory (MBT) Method

For the simple beam theory, the value of strain energy release rate (GI) of a

perfectly built-in is calculated via:

3𝑃𝛿
𝐺𝐼 = (A.1)
2𝑏𝑎
where

P = load,
δ = load point displacement,
b = specimen width, and
a = delamination length.
However, the simple beam theory expression for the compliance of a perfectly

built-in DCB specimen does underestimate the compliance as the beam is not

perfectly built-in. A means of correcting for this effect is to treat the DCB as if
Appendix A 223

it contained a slightly longer delamination, 𝑎 + |∆| , where ∆ may be

determined experimentally by generating a least-squares plot of the cube root

of compliance, C1/3, as a function of delamination length (Figure A.2a). Here,

the compliance C is the ratio of the crack opening displacement to applied load

(δ/P). The values used to generate this plot should be the load and displacements

corresponding to the visually observed delamination onset on the edge and all

the propagation values. So, the GI value is calculated as follows:

3𝑃𝛿
𝐺𝐼 =
2𝑏(𝑎 + |∆|) (A.2)

A.2 Compliance Calibration (CC) Method

An alternative approach to correct the simple beam method is to plot the

logarithm of the compliance C, or of the normalized compliance, C/N, versus

the logarithm of the crack length a as shown in Figure A.2b. It is noteworthy

that only the propagation values are used for the linear fits, and all the initiation

values are not considered in the regression analysis. Calculation of the GI value

is as follows:

3𝑛𝑃𝛿
𝐺𝐼 = (A.3)
2𝑏𝑎
where 𝑛 is the slope of log (𝛿𝑖 /𝑃𝑖 ) versus log (𝑎𝑖 ) line.

A.3 Modified Compliance Calibration (MCC) Method

In this method, the slope (𝐴1 ) of the least squares plot of the delamination length

normalized by specimen thickness, a/h, versus the cube root of compliance, C1/3
224 Calculation of Mode I interlaminar fracture toughness (G IC) using DCB test data

is used to calibrate the beam theory method (Figure A.2c). Calculation of the

GI value is as follows

3𝑃 2 𝐶 2/3
𝐺𝐼 =
2𝐴1 𝑏ℎ (A.4)

(a) (b) (c)

Figure A.2. Definition of correction factors for different data reduction methods; (a) MBT. (b)
CC, (c) MCC [173].
Appendix B

Table B.1. Tensile properties of unstitched and stitched woven flax fiber composites.
composites
Tensile Strength [MPa] Tensile Stiffness [GPa]
Appendix B

Lamina Stitch Stitch Areal Fraction Failure Strain


Laminate
preform material x10-2 [%]
Between 0.1% - 0.3% strain Between 0.1% - 0.3% strain
PW0* Unstitched 0 87.2 (9)† 7.91 (13) 1.62 (3)
PWC1 Cotton 0.46 78.6 (10) 7.82 (12) 1.58 (14)
PWC2 Cotton 0.92 82.1 (10) 8.14 (14) 1.75 (14)
Flax
PWC3 Cotton 1.22 77.3 (11) 7.21 (16) 1.53 (10)
4×4Plain
PWC4 Cotton 1.83 74.3 (12) 7.03 (14) 1.57 (16)
Weave
2 PWF1 Flax 0.69 84.1 (9) 8.02 (12) 1.6 (4)
(500 g/m )
PWF2 Flax 1.03 80.1 (12) 7.87 (10) 1.51 (6)
PWF3 Flax 1.38 75.2 (13) 7.19 (17) 1.47 (6)
PWF4 Flax 2.06 70.5 (15) 6.92 (18) 1.47 (10)
†Values in bracket represent coefficient of variation (CoV%), defined as the ratio of the standard deviation to the average value (in percentage).
* 𝑉𝑓 = 31% for woven flax fiber composites.
Mechanical Properties and Mode I fracture
toughness of unstitched and stitched flax fiber
225
226

Table B.2. Flexural properties of unstitched and stitched woven flax fiber composites.

Flexural Strength [MPa] Flexural Modulus [GPa] Flexural Failure


Lamina Stitch Stitch Areal Fraction
Laminate Strain
preform material x10-2
Between 0.3% - 0.5% strain Between 0.3% - 0.5% strain [%]
PW0* Unstitched 0 125.7 (10)† 5.81 (9) 3.40 (7)
PWC1 Cotton 0.46 116.5 (6) 5.32 (3) 2.88 (7)
PWC2 Cotton 0.92 120.3 (5) 5.72 (7) 2.98 (5)
Flax
PWC3 Cotton 1.22 124.1 (8) 6.01 (6) 3.11 (3)
4×4Plain
PWC4 Cotton 1.83 120.1 (7) 6.22 (9) 2.92 (5)
Weave
2 PWF1 Flax 0.69 123.4 (9) 5.96 (7) 3.04 (4)
(500 g/m )
PWF2 Flax 1.03 121.9 (11) 5.89 (10) 3.01 (5)
PWF3 Flax 1.38 115.6 (12) 5.94 (12) 2.77 (8)
PWF4 Flax 2.06 110.1 (11) 6.06 (10) 2.54 (5)
†Values in bracket represent coefficient of variation (CoV%), defined as the ratio of the standard deviation to the average value (in percentage).
* 𝑉𝑓 = 31% for woven flax fiber composites.
Calculation of Mode I interlaminar fracture toughness (G IC) using DCB test data
Appendix B

Table B.1. Detailed G IC values measured through DCB test of flax and glass fiber laminates.

Stitch Areal GIC – Minimum of initiation [J/m2] GIC – Propagation [J/m2]


Stitch
Lamina preform Laminate Fraction
material
x10-2 MBT CC MCC MBT CC MCC
PW0* Unstitched 0 1827 (8)† 1891 (8) 1931 (8) 3185 (15) 3222 (15) 3226 (15)
PWC1 Cotton 0.46 1763 (9) 1797 (10) 1832 (10) 3224 (19) 3295 (19) 3317 (19)
PWC2 Cotton 0.92 1944 (11) 2008 (12) 2028 (12) 3252 (22) 3305 (22) 3336 (22)
Flax 4×4Plain PWC3 Cotton 1.22 2065 (13) 2098 (13) 2143 (13) 3338 (19) 3368 (19) 3392 (19)
Weave PWC4 Cotton 1.83 2092 (13) 2157 (13) 2170 (13) 3320 (21) 3402 (21) 3444 (21)
(500 g/m2) PWF1 Flax 0.69 2037 (12) 2094 (13) 2127 (13) 3490 (23) 3504 (22) 3597 (22)
PWF2 Flax 1.03 2203 (15) 2297 (15) 2308 (15) 3693 (22) 3708 (22) 3710 (22)
PWF3 Flax 1.38 2394 (21) 2457 (21) 2471 (21) 3713 (23) 3765 (23) 3772 (23)
PWF4 Flax 2.06 2512 (23) 2590 (23) 2612 (23) 3860 (25) 3945 (25) 3960(25)
Flax UD UD0 Unstitched 0 771 (6) 789 (6) 803 (5) 1274 (6) 1308 (6) 1313 (5)
(110 g/m2) UDF Flax 1.38 1344 (12) 1401 (14) 1392 (14) 2223 (13) 2298 (15) 2289 (14)
Glass UD glass Unstitched 0 316 (5) 324 (4) 322 (4) 681 (2) 682 (2) 681 (2)
†Values in bracket represent coefficient of variation (CoV%), defined as the ratio of the standard deviation to the average value (in percentage).
* 𝑉𝑓 = 31% for woven flax composites, 𝑉𝑓 = 40% for UD flax composites, and 𝑉𝑓 = 61% for UD glass fiber composite.
227
228 Testing for shear damage parameter of woven flax fiber composite (d12)

Appendix C
Testing for shear damage parameter of woven
flax fiber composite (d12)

This appendix presents the experimental procedure for determining the shear

damage parameter of the woven flax fabric composite which is used in the FEM

analysis of this study to model the in-plane material behavior of matrix. This

method follows the procedure proposed by Johnson [158] and ASTM D3518

[223].

As discussed previously in Chapter 7, it is assumed that failure in principle fiber

directions (tension and compression) and in-plane shear are decoupled. The

elastic constants (E11, E22, v21) and strength of the material along the fiber

directions under tension and compression ( 𝑋𝑇± , 𝑋𝐶± ) loading can be easily

measured by the standard uniaxial tensile/compression test of composite in fiber

directions, where the fiber failure is the predominant failure mode. The damage

evolution in the fiber failure modes is accomplished based on the corresponding

fracture toughness which may be determined experimentally

(tension/compression failure of composite along the fiber directions).

The in-plane shear response of the laminate is characterized by means of a cyclic

tensile loading at 45° to the principle fiber directions, since the monotonic

loading is not sufficient to determine all the parameters needed to model the

elastic-plastic response and damage evolution. Here, it is assumed that the

effects of strain along the fiber directions on the shear response is negligible.

Figure C.1 shows the tensile test on a ±45 layup of woven flax fiber laminate.
Appendix C 229

𝜎12

𝜎12

-45 ° +45°

Figure C.1. Tensile test of the woven flax fiber laminate with ±45 layup to measure the shear
response.

The shear response of the woven flax reinforced composite under monotonic

and cyclic loading, which was obtained with six load-unload cycles, are

compared in Figure C.2. The in-plane shear response (stress and strain) was

measured in accordance with test method D3518 [223]. The slight difference

between the stress levels of the monotonic and cyclic test can be attributed to

stress relaxation due to material plasticity. In what follows, the result of cyclic

loading is used to calibrate the parameters of the shear damage and plasticity

model.
230 Testing for shear damage parameter of woven flax fiber composite (d12)

Monotonic loading C4 C5 C6
50
Cyclic loading C3

40 C2
Shear stress, σ12 (MPa)

30
C1

20

10 (1 − 𝑑12 )𝐺12

0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Shear strain, ε12 (%)
𝑝𝑙 𝑒𝑙
𝜀12 𝜀12

Figure C.2. Measured monotonic and cyclic shear stress-strain curves for Woven flax/epoxy
laminate.

As can be seen from Figure C.2, the total strain 𝜀12 at each cycle can be written

𝑒𝑙 𝑝𝑙
as the sum of an elastic 𝜀12 and plastic 𝜀12 strain component, given as:

𝑒𝑙 𝑝𝑙
𝜀12 = 𝜀12 + 𝜀12
(C.1)

The elastic shear strain component can be calculate (form the unloading curve)

using the linear elastic concept as

𝑒𝑙
𝜎12
𝜀12 = 𝑠𝑒𝑐
𝐺12
(C.2)
𝑠𝑒𝑐
in which 𝐺12 is the secant shear modulus at each unloading cycle. The shear

damage parameter 𝑑12 can be measured form the ratio of the secant (damaged)

shear modulus to the undamaged (initial) elastic modulus 𝐺12 as


Appendix C 231

𝑠𝑒𝑐
𝐺12 = 𝐺12 (1 − 𝑑12 )
(C.3)

The shear damage parameter 𝑑12 can be determined from the measured applied

stress level and elastic shear strain using Eq. (C.2) and Eq. (C.3). Thus, the stress

- damage values (𝜎12 , 𝑑12 ) pairs can be calculated for each (linear) unloading

curve. However, representing the damager parameter against the stress values

is not a proper basis to provide an accurate linear model of test data. A simple

function of the stress which gives a reasonably accurate linear model for the

damage parameter is a logarithmic function of stress, ln(𝜎̃12 ), where 𝜎̃12 =

𝜎12 /(1 − 𝑑12 ) is the effective linear elastic stress. Figure A.3 shows the shear

damage parameter against ln(𝜎̃12 ) measured at each loading cycle and a linear

fit to the test data. The damage evolution constant 𝛼12 and the shear yield stress

threshold 𝑆 are obtained from the gradient of the fitted line to the data and its

intersection with the horizontal axis, respectively.

0.7
𝑚𝑎𝑥
𝑑12 ≤1
0.6

0.5

0.4
d12

0.3
𝛼12
0.2

0.1

0
2.7 3.7 4.7 5.7 6.7

ln( 12)

Figure C.3. Elastic shear damage evolution against 𝑙𝑛 (𝜎̃12 ) for calibration of the damage
parameters 𝛼12 and 𝑆 for the woven flax fiber laminate.
232 Testing for shear damage parameter of woven flax fiber composite (d12)

To determine the plastic hardening function under shear deformation, the

𝑝𝑙
accumulated plastic strain has to be determined. The plastic strain 𝜀12

attributable to each loading cycle, is obtained from the residual deformation

𝑝𝑙
after each unloading cycle (Figure C.2). The values of 𝜀12 are then used to plot

against the corresponding effective shear stress 𝜎̃12 at the onset of unloading to

determine the shear hardening function, as shown in Figure C.4 As can be seen,

𝑝𝑙 𝐾
the test data point are well fitted by the power law function, 𝜎̃𝑦0 + 𝐶(𝜀12 ) ,

with calibrated plasticity parameters of 𝐶 and 𝐾. The value of the hardening

function at a zero plastic strain (intersection with the vertical axis) is the

inelastic threshold shear stress 𝜎̃𝑦0 .

140

120

100
̃12 (Mpa)

𝐾
𝑝𝑙
80 𝜎̃𝑦0 + 𝐶 𝜀12

60
σ

40

20 Experimental
Series1

0
0 0.5 1 1.5 2 2.5
l
12 ( )

Figure C.4. Plastic hardening function in shear for the woven flax fiber composite.
Appendix C 233

The shear elastic properties and aforementioned damage parameters obtained

from the cyclic shear test of the woven flax fiber laminate are summarized in

Table C.1.

Table C.1. Shear elastic properties and calibrated shear damage constants of the woven flax
fiber composite.

𝜎̃𝑦0 (𝑀𝑃𝑎) 𝑚𝑎𝑥


𝐺12 (𝐺𝑃𝑎) 𝑆 (𝑀𝑃𝑎) 𝑑12 𝛼12 𝐶 𝐾
3.97 19.71 23.9 0.63 0.32 68 0.45

Vous aimerez peut-être aussi