Liu-J-2021-PhD-Thesis (4)

Download as pdf or txt
Download as pdf or txt
You are on page 1of 183

THE PERFORMANCE OF THERMOPLASTIC

AND THERMOSET COMPOSITES SUBJECTED


TO IMPACT LOADING
by

Jun Liu

A thesis submitted for the degree of Doctor of Philosophy of Imperial College


London and the Diploma of Imperial College London

Department of Mechanical Engineering

Imperial College London

December 2020
Abstract

Carbon Fibre Reinforced Polymer (CFRP) composites play an increasingly important


role in the aerospace industry where CFRPs are extensively employed in the primary structures
of aircrafts. These composite materials have demonstrated relatively high stiffness-to-weight
and strength-to-weight ratios, as well as good fatigue and corrosion resistance, compared with
the conventional metallic materials. However, these composite primary structures of aircrafts
face great challenges when encounter with an impact with foreign bodies. Due to the
anisotropic property and interaction between fibres and matrix, the failure of composite
structure is generally a complex process, which requires a further investigation and better
understanding.

Bearing the above in mind, the aim of this project is to investigate the modes and extent
of damage inflicted on the unidirectional carbon fibre and woven carbon fibre reinforced
thermoplastic and thermoset composites, under various impact loadings. A series of low-
velocity (drop-weight) and high-velocity (gas-gun) experiments were firstly performed, using
rigid impactors at different energy levels, on the unidirectional Carbon Fibre Reinforced Poly
(ether-ether-ketone) (CF/PEEK) and toughened Epoxy (CF/Epoxy) cross-ply composites.
Particularly, for the high-velocity experiments, the three-dimensional (3D) Digital Image
Correlation (DIC), coupling with high-speed cameras, is employed to measure the out-of-plane
displacement from the rear face of the composite specimens during the impact events. The
ultrasonic C-scan results are utilised for assessing damage in the composites and reflecting the
strain rate effects on the impact response of the composites. A finite element model, based on
the continuum shell elements, is developed to simulate both the low- and high-velocity impact
test results. The modelling results are quantitatively validated against the experimentally
obtained impact behaviour of the composites, such as the loading response, interlaminar and
intralaminar damage etc. The experimental validation enables the model to be used with
confidence in the future industrial applications.

I
To better understand the failure mechanisms of woven-fabric composites, woven-fabric
CF/PEEK and CF/Epoxy specimens are employed to study the behaviour of woven composite
laminates subjected to impact loading by soft gelatine and hard metallic projectiles to represent
the impact of foreign object, such as a small bird, hail-stone or runway debris. For this
investigation, both the 3D DIC measurements and C-scan assessments are employed in the
experiments performed on the CF/PEEK and CF/Epoxy specimens. A FE model is developed,
based on Smoothed Particle Hydrodynamics (SPH) method, Abaqus built-in Hashin’s criteria
and bilinear cohesive law, to predict the viscoelastic-plastic fluid behaviour of the gelatine
project and the behaviour of composite specimens during the impact events. Good agreement
is shown between the predictions from using the FE model and the experimental results. The
FE model has produced accurate predictions of the deformation of the gelatine projectiles, the
major strain and out-of-plane displacement of the composites, contact pressure and the damage
degrees resulting in the rear-face of the composite specimens. The dynamic characteristics of
the thermoplastic and thermoset composite laminates, investigated in the project through a
series of experimental tests and numerical predictions, can assist the design of lightweight
composite structures for energy-absorbing applications.

II
Declaration of Originality

I hereby affirm that the work presented in this thesis was produced by me during the
course of my Ph.D. studies at Imperial College London. Where contributions by others are used,
they have been appropriately referenced in the test and in the bibliography. No part of the work
presented in this thesis has been submitted elsewhere for another degree or qualification.

Copyright Declaration

The copyright of this thesis rests with the author. Unless otherwise indicated, its
contents are licensed under a Creative Commons Attribution-Non Commercial 4.0
International Licence (CC BY-NC). Under this licence, you may copy and redistribute the
material in any medium or format. You may also create and distribute modified versions of the
work. This is on the condition that: you credit the author and do not use it, or any derivative
works, for a commercial purpose. When reusing or sharing this work, ensure you make the
licence terms clear to others by naming the licence and linking to the licence text. Where a
work has been adapted, you should indicate that the work has been changed and describe those
changes. Please seek permission from the copyright holder for uses of this work that are not
included in this licence or permitted under UK Copyright Law.

III
Acknowledgements

I would like to express my greatest respect and the deepest gratitude to my supervisor
Professor John P. Dear. His constant pursuit of details and perfection, experienced perspective
of the overall aim had a great positive influence on my professional and personal development.
Without his strong support and excellent guidance, I would not have the honour to accomplish
this study. Also, I would like to express my sincere gratitude to my co-supervisors Dr. Haibao
Liu, Professor Bamber R.K. Blackman, Professor Anthony J. Kinloch FREng FRS and
Professor Roderick A. Smith FREng. Their constructive ideas and considerably experience
always guide me to achieve better results.

Particularly, I would like to express my thanks to my colleagues and technicians at the


Imperial College London, especially in our Mechanics of Materials Division. Many thanks to
Professor Jin Zhou, Dr. Xiangshao Kong, Mr. Wengui Wang, Ms. Jie Zheng, Dr. Hui Chai and
Mr. Mengen Xing for their valuable advice. Many thanks to Mr. Yuzhe Ding, Ms. Zoe Hall
and Dr. Cihan Kaboglu for their assistance with experimental testing. Special thanks to Dr.
Zhaoheng Cai and Dr. Xiao Hua for their support through my study.

The strong support from the Aviation Industry Corporation of China (AVIC), the First
Aircraft Institute (FAI) and the Manufacturing Technology Institute (MTI) for funding this
research is much appreciated.

At last, I would like to thank my family, especially my mother, for all the support that
have shown me since my primary school. My special thanks to my wife Mingyan and my son
Hongyi, without whom I would not have enough confidence and patience to complete the
studies.

IV
Contents

List of Figures .......................................................................................................................... X

List of Tables ...................................................................................................................... XVI

Nomenclature .................................................................................................................. XVIII

Abbreviation ....................................................................................................................... XXI

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

1.1 Background ................................................................................................................. 1

1.2 Aims and objectives .................................................................................................... 4

1.3 Structure of the thesis .................................................................................................. 4

Chapter 2. Literature review ............................................................................................... 7

2.1 Introduction ................................................................................................................. 7

2.2 Composite materials .................................................................................................... 7

2.2.1 Classification of composite materials .................................................................. 7

2.2.2 Fibre reinforcement .............................................................................................. 8

2.2.3 Polymer matrix................................................................................................... 11

2.3 Fabrication processes and their effects...................................................................... 13

2.3.1 Fabrication processes ......................................................................................... 13

2.3.2 Effects of manufacturing process....................................................................... 15

2.4 Impact performance................................................................................................... 16

2.4.1 Under low-velocity impact loading ................................................................... 17

2.4.2 Under high-velocity impact loading .................................................................. 20

2.4.3 Typical failure modes ........................................................................................ 24

V
CONTENTS

2.5 Techniques for detecting impact damage in fibre reinforced polymer composites .. 27

2.5.1 Ultrasonic C-scanning ........................................................................................ 27

2.5.2 X-ray Computed Tomography (CT) scanning ................................................... 29

2.5.3 Other techniques ................................................................................................ 30

2.6 Composite damage modelling ................................................................................... 31

2.6.1 Interlaminar modelling....................................................................................... 31

2.6.2 Intralaminar modelling....................................................................................... 32

2.6.3 Modelling impact behaviour .............................................................................. 33

2.7 Summary ................................................................................................................... 35

Chapter 3. Materials and Experimental procedures ....................................................... 37

3.1 Introduction ............................................................................................................... 37

3.2 Materials .................................................................................................................... 37

3.2.1 Unidirectional fibre reinforced composites ....................................................... 37

3.2.2 Woven-fabric fibre reinforced composites ........................................................ 39

3.3 Manufacturing processes ........................................................................................... 40

3.3.1 Thermoplastic polymer composites ................................................................... 40

3.3.2 Thermoset polymer composites ......................................................................... 42

3.4 Overview of experiments .......................................................................................... 44

3.4.1 Low-velocity (drop weight) testing ................................................................... 44

3.4.2 High-velocity (gas gun) testing .......................................................................... 45

3.4.3 3D Digital Image Correlation (DIC) .................................................................. 46

3.4.4 Ultrasonic C-scan ............................................................................................... 48

Chapter 4. Experimental study on unidirectional thermoplastic and thermoset


laminates under low-velocity and high-velocity impact loading........................................ 49

4.1 Introduction ............................................................................................................... 49

4.2 Test specification and methods ................................................................................. 50

4.2.1 Drop weight testing ............................................................................................ 50

VI
CONTENTS

4.2.2 Gas gun testing ................................................................................................... 50

4.2.3 Post-test inspection ............................................................................................ 51

4.3 Experimental results .................................................................................................. 51

4.3.1 Low-velocity drop weight tests .......................................................................... 51

4.3.2 High-velocity gas gun impact tests .................................................................... 57

4.3.3 Comparison of damage inflicted by the low-velocity and high-velocity tests... 61

4.4 Summary ................................................................................................................... 62

Chapter 5. Numerical modelling on unidirectional thermoplastic and thermoset


laminates under low-velocity and high-velocity impact loading........................................ 64

5.1 Introduction ............................................................................................................... 64

5.2 Composite damage model ......................................................................................... 64

5.2.1 Brief overview of the damage model ................................................................. 64

5.2.2 The intralaminar damage model ........................................................................ 67

5.2.3 The interlaminar damage model ........................................................................ 71

5.3 Modelling results ....................................................................................................... 74

5.3.1 Material properties ............................................................................................. 74

5.3.2 Simulations of the low-velocity drop weight impact test results ....................... 74

5.3.3 Simulations of the high-velocity gas gun impact test results ............................. 80

5.4 Summary ................................................................................................................... 86

Chapter 6. Experimental study on woven-fabric thermoplastic and thermoset


laminates under high-velocity soft and hard impact loading............................................. 87

6.1 Introduction ............................................................................................................... 87

6.2 Test specification and methods ................................................................................. 88

6.2.1 Test specimens and projectiles........................................................................... 88

6.2.2 Gas gun testing ................................................................................................... 90

6.2.3 Post-test inspection ............................................................................................ 92

6.3 Experimental results .................................................................................................. 93

6.3.1 The soft projectile impact testing ....................................................................... 93


VII
CONTENTS

6.3.2 The hard projectile impact testing...................................................................... 99

6.4 Summary ................................................................................................................. 102

Chapter 7. Experimental study and numerical modelling on woven-fabric


thermoplastic and thermoset laminates under high-velocity impact loading ................ 104

7.1 Introduction ............................................................................................................. 104

7.2 Numerical modelling methods ................................................................................ 105

7.2.1 Modelling the response of the projectile .......................................................... 105

7.2.2 Modelling the response of the composites ....................................................... 107

7.3 Experimental test specification and methods .......................................................... 108

7.3.1 Projectiles and composites ............................................................................... 108

7.3.2 Experimental investigations ............................................................................. 108

7.4 Experimental results ................................................................................................ 109

7.4.1 Deformation of the gelatine projectile ............................................................. 109

7.4.2 Effects of the impact energy of the gelatine projectile .................................... 110

7.4.3 Effects of the matrix system ............................................................................ 113

7.5 The Finite Element (FE) model ............................................................................... 115

7.5.1 Model definition............................................................................................... 115

7.5.2 Input parameters............................................................................................... 116

7.5.3 Implementation of the model ........................................................................... 119

7.6 Model validation and application ............................................................................ 119

7.6.1 Validation of the model ................................................................................... 119

7.6.2 Application of the model ................................................................................. 122

7.7 Summary ................................................................................................................. 126

Chapter 8. Conclusions ..................................................................................................... 128

8.1 Introduction ............................................................................................................. 128

8.2 Effects of the matrix system employed on the performance of CFRPs .................. 128

8.3 Effects of the loading rate on the impact performance of CFRPs ........................... 129

VIII
CONTENTS

8.4 Effects of the hardness of the projectile on the impact response of CFRPs............ 130

8.5 Modelling unidirectional CFRPs under low-velocity and high-velocity impact .... 130

8.6 Modelling woven-fabric CFRPs under high-velocity impact ................................. 131

Chapter 9. Future work .................................................................................................... 132

9.1 Introduction ............................................................................................................. 132

9.2 Effects of impact loading on compression-after-impact (CAI) strength of CFRPs 132

9.3 Effects of environmental factors on impact performance of CFRPs ...................... 133

9.4 Performance of the repaired CFRPs under impact loadings ................................... 133

List of publication ................................................................................................................ 135

Bibliography ......................................................................................................................... 137

Appendices ............................................................................................................................ 154

Appendix A: Effects of impactor geometry on the low-velocity impact behaviour of fibre-


reinforced composites: an experimental investigation ..................................... 154

Appendix B: Summary of copyright permissions .................................................................. 160

IX
List of Figures

Figure 1-1. Applications of composite materials: (a) Bombardier Transportation C20 metro
train with composite car body; (b) a Formula One car with carbon fibre chassis; (c)
Airbus A350-XWB aircraft employed 53 wt% of composite materials [8–10]. ............... 1
Figure 1-2. Composites in aerospace industry: (a) composite structural weight development
on Airbus aircrafts; (b) composite solutions applied throughout the Boeing Dreamliner
787 [10,11]. ........................................................................................................................ 2
Figure 1-3. Aircraft structures collapsed after (a) hail impact, or (b) bird-strike [17,18]. ........ 3
Figure 2-1. Typical structure of composite [20]. ....................................................................... 8
Figure 2-2. The diagram of tensile stress-strain for various reinforcing fibres [3].................... 9
Figure 2-3. Illustration of fibre alignment in two layers of a cross-ply UD prepreg laminate
[26]. .................................................................................................................................. 11
Figure 2-4. Molecular structure of thermoplastic and thermoset polymers [28]. .................... 11
Figure 2-5. Schematic of a vacuum bag system [4]. ................................................................ 14
Figure 2-6. Schematic of a compression moulding system [3]. ............................................... 15
Figure 2-7. Schematic representation of impact response under low and high-velocity impact
[46]. .................................................................................................................................. 17
Figure 2-8. Schematic of drop-weight impact fixture in the ASTM 7136 standard [47]. ....... 18
Figure 2-9. Illustration of impact damage in a laminate [55]. ................................................. 21
Figure 2-10. The main failure types shown in the unidirectional FRP laminates: (a)
experiemtnal observed failure modes [43] and (b) schematically categorised failure
modes. .............................................................................................................................. 24
Figure 2-11. Modes of interlaminar failure [81]. ..................................................................... 25
Figure 2-12. Fibre-dominated (a) tensile and (b) compressive failure modes [90]. ................ 25
Figure 2-13. Fracture planes obtained from woven CFRPs under (a) tensile and (b)
compressive loading [91]. ................................................................................................ 26

X
LIST OF FIGURES

Figure 2-14. Shear stress-strain curve for the static [+45°, -45°]2s composite specimens under
tensile tests [97]. .............................................................................................................. 27
Figure 2-15. Representations of (a) A-scan data and (b) C-scan data [22].............................. 28
Figure 2-16. Stresses and fracture plane on a composite material [125]. ................................ 33
Figure 2-17. Out-of-plane displacement magnitude for stiffened composite panel FE model
after impact showing permanent indentation as a result of the damage process (in mm).
.......................................................................................................................................... 34
Figure 3-1. Photographs of the unidirectional CF/PEEK and CF/epoxy composite specimens.
.......................................................................................................................................... 38
Figure 3-2. Photographs of the woven-fabric CF/PEEK and CF/Epoxy composite specimens.
.......................................................................................................................................... 39
Figure 3-3. Photograph of (a) the thermo compressor and the diagram of (b) the consolidation
jig and (c) the cure cycle for the CF/PEEK laminates consolidation. ............................. 41
Figure 3-4. Photograph of (a) the autoclave and the diagram of (b) the consolidation jig and
(c) the cure cycle for the CF/Epoxy laminates consolidation. ......................................... 43
Figure 3-5. Set-up for the drop weight (low-velocity) experiments. ....................................... 45
Figure 3-6. Set-up for the gas gun (high-velocity) experiments. ............................................. 46
Figure 3-7. Schematic for the high-velocity impact test with 3D DIC system. ....................... 47
Figure 3-8. Portable C-scan system. ........................................................................................ 48
Figure 4-1. Measured load as a function of time curves for the low-velocity drop-weight tests
at impact energies (IE) of 4.5, 7.5 and 10.5 J: (a) CF/PEEK, (b) CF/Epoxy and (c) a
comparison of typical load versus time curves for the CF/PEEK and CF/Epoxy
composites........................................................................................................................ 52
Figure 4-2. Measured load as a function of displacement curves for the low-velocity drop-
weight tests for the CF/PEEK and CF/Epoxy specimens at impact energies of (a) 4.5 J,
(b) 7.5 J and (c) 10.5 J...................................................................................................... 54
Figure 4-3. C-scan damage maps obtained from the (a) CF/PEEK and (b) CF/Epoxy replicate
composite test specimens impacted at a low-velocity (i.e. the drop-weight test) at impact
energies of 4.5 J, 7.5 J and 10.5 J. The corresponding impact velocities were 1.68, 2.16
and 2.56 m.s-1, respectively.............................................................................................. 55
Figure 4-4. C-scan damage maps obtained from the CF/PEEK and CF/Epoxy duplicate
composite test specimens impacted at a high-velocity (i.e. the gas-gun test). The average
impact velocity and energy were 54.4±1.0 m.s-1 and 10.5±0.3 J, respectively. .............. 57

XI
LIST OF FIGURES

Figure 4-5. Maps and diagrams of the typical full-field out-of-plane displacements of the rear
surface of the composite specimens measured for both the loading (upper map strips)
and unloading (lower map strips) phases for the high-velocity gas-gun experiments at
54.4±1.0 m.s-1 and an impact energy of 10.5±0.3 J for the CF/PEEK composite test
specimens. ........................................................................................................................ 59
Figure 4-6. Maps and diagrams of the typical full-field out-of-plane displacements of the rear
surface of the composite specimens measured for both the loading (upper map strips)
and unloading (lower map strips) phases for the high-velocity gas-gun experiments at
54.4±1.0 m.s-1 and an impact energy of 10.5±0.3 J for the CF/Epoxy composite test
specimens. ........................................................................................................................ 60
Figure 5-1. The implementation of the FE model showing schematically the flow chart, for
one computation time-step, for a single element for modelling the interlaminar and
intralaminar damage......................................................................................................... 65
Figure 5-2. Schematic of the FE model. .................................................................................. 66
Figure 5-3. Schematic of the bilinear cohesive surface law. ................................................... 72
Figure 5-4. Experimental and simulated (red-coloured) footprints of the damage area for the
CF/PEEK composite specimens for impact energies of 4.5, 7.5 and 10.5 J for the low-
velocity drop-weight impact test. ..................................................................................... 76
Figure 5-5. Experimental and simulated (red-coloured) footprints of the damage area for the
CF/Epoxy composite specimens for impact energies of 4.5, 7.5 and 10.5 J for the low-
velocity drop-weight impact test. ..................................................................................... 76
Figure 5-6. Simulations of the (a) interlaminar delamination (red-coloured) damage and (b)
intralaminar matrix (red-coloured) damage as a function of time-scale for the CF/PEEK
composites for low-velocity drop-weight tests at an impact energy and velocity of 10.5 J
and 2.56 m.s-1, respectively.............................................................................................. 78
Figure 5-7. Experimental and simulated load versus time curves for the low-velocity drop-
weight tests for the CF/PEEK composite specimens for: (a) 4.5, (b) 7.5 and (c) 10.5 J
impact energy levels. ....................................................................................................... 79
Figure 5-8. Experimental and simulated load versus time curves for the low-velocity drop-
weight tests for the CF/Epoxy composite specimens for: (a) 4.5, (b) 7.5 and (c) 10.5 J
impact energy levels. ....................................................................................................... 79
Figure 5-9. Experimental and simulated (red-coloured) footprints of the damage areas for the
CF/PEEK and CF/Epoxy composite specimens from the high-velocity gas-gun test using
an impact energy and velocity of 10.5±0.3 J and 54.4±1.0 m.s-1, respectively. .............. 82
XII
LIST OF FIGURES

Figure 5-10. The experimental and simulated out-of-plane displacement maps obtained from
(a) the CF/PEEK and (b) the CF/Epoxy composite specimens when subjected to the
high-velocity gas-gun test at an impact energy and velocity of 10.5±0.3 J and 54.4±1.0
m.s-1, respectively. ........................................................................................................... 84
Figure 5-11. Load versus time curves as predicted from the FE model for the CF/PEEK and
CF/Epoxy composite specimens subjected to a high-velocity gas-gun test at an impact
energy and velocity 10.5 J and 54.4 m.s-1, respectively.. ................................................ 85
Figure 6-1. Projectiles and sabot for soft- and hard-impact tests: (a) flat-fronted soft gelatine
projectile and (b) HDPE sabot and (c) hemi-spherical head of the hard aluminium-alloy
projectile.. ........................................................................................................................ 88
Figure 6-2. Schematic of the support for the composite test specimen. .................................. 91
Figure 6-3. Set-up II for the high-velocity hard-projectile impact: (a) photograph and (b)
schematic.......................................................................................................................... 92
Figure 6-4. The cut-off machine equipped with high concentration diamond saw. ................ 93
Figure 6-5. Experimental results for the CF/PEEK panel impacted using a soft-gelatine
projectile with a velocity of 100 m.s-1 showing the (a) out-of-plane displacement maps
and (b) out-of-plane displacement profile during the loading and unloading phases. ..... 94
Figure 6-6. Experimental results for the CF/Epoxy panel impacted using a soft-gelatine
projectile with a velocity of 100 m.s-1 showing the (a) out-of-plane displacement maps
and (b) out-of-plane displacement profile during the loading and unloading phases. ..... 95
Figure 6-7. Maximum out-of-plane displacement versus the impact energy for the CF/PEEK
and CF/Epoxy composites upon being impacted using soft-gelatine projectiles. ........... 96
Figure 6-8. Visual inspection of the rear surface of the (a) CF/PEEK specimens and the (b)
CF/Epoxy specimens subjected to different impact velocities by the soft-gelatine
projectile. ......................................................................................................................... 97
Figure 6-9. Photographs of the cross-section of the mid-plane of the composite: (a) CF/PEEK
and (b) CF/Epoxy specimens impacted using a soft-gelatine projectile with an impact
velocity of 100 m.s-1. ....................................................................................................... 98
Figure 6-10. Ultrasonic C-scan maps of the (a) CF/PEEK specimens and (b) CF/Epoxy
specimens subjected to impact velocities using the soft-gelatine projectile and (c)
magnified C-scan image of the CF/Epoxy specimen impacted at the highest velocity of
100 m.s-1........................................................................................................................... 99

XIII
LIST OF FIGURES

Figure 6-11. Photographs of the rear surface of the (a) CF/PEEK and (b) CF/Epoxy
composites impacted using the hard aluminium-alloy projectile at various initial impact
velocities. ....................................................................................................................... 100
Figure 6-12. Comparison of the kinetic energy absorption (KEA) and damage area (DA)
versus the impact energy for the CF/PEEK and CF/Epoxy composites impacted using
the hard aluminium-alloy projectile at different velocities. ........................................... 101
Figure 7-1. Schematics of the types of post-impact damage on the rear-face of the
composites: (a) no visible damage, (b) cracking, (c) fracture and (d) perforation. ....... 109
Figure 7-2. Deformation history of the gelatine projectile for a 37 J impact energy impacting
the CF/PEEK composite. ............................................................................................... 110
Figure 7-3. CF/PEEK composites impacted at a 37 J energy level: (a) the OOP displacement
contours and (b) the evolution of the OOP displacement profiles (in intervals of 0.025
ms) during loading and unloading. ................................................................................ 112
Figure 7-4. Photographs of the rear-faces of the CF/PEEK composites after impact: (a) for an
energy of 37 J (‘Test GCP-I’) and (b) for an energy of 72 J (‘Test GCP-IV’). ............. 113
Figure 7-5. The rear-faces of the specimens after impact: (a) the CF/PEEK composite
impacted at a 37 J energy level and (b) the CF/epoxy composite impacted at a 38 J
energy level. ................................................................................................................... 114
Figure 7-6. The FE model with PC3D elements. ................................................................... 115
Figure 7-7. The creation of a single equivalent [0o-90o] woven-fibre reinforced composite ply.
........................................................................................................................................ 116
Figure 7-8. Deformation of the gelatine projectile obtained from the experimental studies and
as predicted from the numerical FE model for the CF/PEEK composites at an impact
energy of 37 J. ................................................................................................................ 120
Figure 7-9. Comparison between the predicted and experimental results for the CF/PEEK
composite at an impact energy of 37 J: (a) the maximum major strain and the out-of-
plane (OOP) displacement and (b) the central OOP displacement versus time trace. ... 121
Figure 7-10. The experimentally measured and predicted degrees of damage resulting in the
rear-face of the CF/PEEK composites at impact energies of (a) 37 J and (b) 72 J. ....... 122
Figure 7-11. Predicted (a) central out-of-plane (OOP) displacement versus time trace and (b)
the experimentally measured and predicted maximum OOP displacements for the
CF/PEEK impacted at 37 J and the CF/epoxy impacted at 38 J. ................................... 123

XIV
LIST OF FIGURES

Figure 7-12. Comparison of the damage obtained from the experiments and the FE
modelling: (a) the CF/PEEK composite impacted at 37 J and (b) CF/epoxy composite
impacted at 38 J. ............................................................................................................ 125
Figure 7-13. Numerical predictions from the FE model for: (a) the average contact pressure
versus time history and (b) the maximum average contact pressure. For the CF/PEEK
composite impacted at an energy of 37 J and (b) the CF/epoxy impacted an energy of 38
J. ..................................................................................................................................... 126

XV
List of Tables

Table 2-1. The comparison of properties between thermoplastics and thermosets [21,22,29].
.......................................................................................................................................... 13
Table 3-1. Primary properties of the unidirectional CF/PEEK and CF/Epoxy laminates. ...... 39
Table 3-2. Primary properties of the woven-fabric CF/PEEK and CF/Epoxy laminates. ....... 40
Table 4-1. Comparison of the measured damage initiation load (i.e. the first major drop in the
measured load versus time curve) and the maximum load obtained from the low-velocity
drop-weight tests.. ............................................................................................................ 53
Table 4-2. Comparison of the footprint of the damage areas obtained from the low-velocity
drop-weight tests as measured from the C-scans. ............................................................ 56
Table 4-3. Comparison of the footprint of the damage areas as measured from the C-scans
obtained from the low-velocity drop-weight and high-velocity gas-gun tests at a similar
impact energy of 10.5±0.3 J. ............................................................................................ 61
Table 5-1. Input properties for the FE modelling studies of the CF/PEEK and CF/Epoxy
composites [133,143,149,155,162–166]. ......................................................................... 75
Table 5-2. Comparison of the experimental and numerically-simulated results obtained from
the low-velocity drop-weight tests at an impact energy and velocity of 10.5 J and 2.56
m.s-1, respectively. ........................................................................................................... 80
Table 6-1. Detailed preparation procedure of the gelatine projectiles. .................................... 90
Table 6-2. Results of the high-velocity soft-gelatine projectile impact tests. .......................... 96
Table 6-3. Results of the high-velocity hard aluminium-alloy projectile impact tests. ......... 101
Table 7-1. Test configurations for investing the effects of the impact velocity and energy on
the CF/PEEK composites. .............................................................................................. 110
Table 7-2. Main DIC results for the CF/PEEK composites impacted by the gelatine
projectiles. ...................................................................................................................... 111
Table 7-3. Gas-gun test conditions to study the effect of the matrix system. ........................ 114

XVI
LIST OF TABLES

Table 7-4. Results from the CF/PEEK and CF/Epoxy composites impacted by the gelatine
projectiles. ...................................................................................................................... 114
Table 7-5. Input properties for the FE modelling of the soft-gelatine projectile [37-39]. ..... 117
Table 7-6. Input properties for the FE modelling studies of the composite
[133,143,149,155,162–165]. .......................................................................................... 118
Table 7-7. Comparison of the experimentally measured and the numerically predicted
maximum central out-of-plane (OOP) displacement. .................................................... 124

XVII
Nomenclature

d General form of the damage variable


𝒅𝒇 Damage variable in fibre damage
𝒅𝒕𝒇 Damage parameter in tensile fibre failure
𝒅𝒄𝒇 Damage parameter in compressive fibre failure
𝒅𝒎 Damage variable in matrix damage
𝒅𝒕𝒎 Damage parameter in tensile matrix failure
𝒅𝒄𝒎 Damage parameter in compressive matrix failure
𝒅𝒔 Damage variable in shear damage
k Stiffness of cohesive law
m Mass of projectile
p Positive pressure in compression
𝒑𝑯 The Hugoniot pressure
s Slope of the linear relationship between 𝑈𝑠 and 𝑈𝑝
t Time
𝒕𝒆 Thickness of an adjacent composite ply
𝒕𝟎𝒊 Initial cohesive law strengths (𝑖 = 33,31,32)
𝒗𝒊 Initial impact velocity
𝒗𝒓 Residual velocity
𝑪𝒅 Damaged elasticity matrix
𝑬𝒊𝒊 Elastic modulus in the fibre or transverse directions (𝑖 = 1,2)
𝑬𝒂 Kinetic energy absorption
𝑬𝒎 Internal energy per unit mass
𝑬𝑯 Specific energy per unit mass
𝑭𝒕𝒇 Tensile damage criterion in fibre failure

XVIII
NOMENCLATURE

𝑭𝒄𝒇 Compressive damage criterion in fibre failure


𝑭𝒕𝒎 Tensile damage criterion in matrix failure
𝑭𝒄𝒎 Compressive damage criterion in matrix failure
G12 Shear modulus in the longitudinal-transverse direction
𝑮𝒄 Interlaminar fracture energy
𝑮𝑰 Mode I energy-release rate
𝑮𝑰𝒄 Interlaminar Mode I fracture energy
𝑮𝑰𝒄 |𝒇𝒕 Tensile intralaminar fracture energy in the longitudinal direction
𝑮𝑰𝒄 |𝒇𝒄 Compressive intralaminar fracture energy in the longitudinal direction
𝑮𝑰𝒄 |𝒎𝒕 Tensile intralaminar fracture energy in the transverse direction
𝑮𝑰𝒄 |𝒎𝒄 Compressive intralaminar fracture energy in the transverse direction
𝑮𝑰𝑰 Mode II energy-release rate
𝑮𝑰𝑰𝒄 Interlaminar Mode II fracture energy
𝑺𝒅 Damage area
𝑺𝑻 Shear strength in the transverse direction
𝑺𝑳 Shear strength in the longitudinal direction
𝑼𝒑 Particle velocity of the projectile
𝑼𝒔 Shock-wave velocity
𝑿𝑻 Tensile strength in the longitudinal direction
𝑿𝑪 Compressive strength in the longitudinal direction
𝒀𝑻 Tensile strength in the transverse direction
𝒀𝑪 Compressive strength in the transverse direction
𝜶 A constant in the cohesive law
 Stress
𝒏 Stress in the normal direction of the composite ply
𝟎𝒏 Strength in the normal direction of the composite ply
𝝈
̂ Effective stress in the composite ply
𝝈
̂ 𝒊𝒊 Effective stress in the longitudinal or transverse directions (𝑖 = 1,2)
𝒔 Shear strength of the composite ply
𝟎𝒔 Shear strength of the cohesive surface
𝒕 Tensile strength of the composite ply
𝟎𝒕 Tensile strength of the cohesive surface

XIX
NOMENCLATURE

𝝉̂𝟏𝟐 Effective stress tensor in the longitudinal-transverse shear direction


𝝂𝒊𝒋 The Poisson’s ratio ( 𝑖, 𝑗 = 1,2, 𝑖 ≠ 𝑗)
𝜺 Strain
𝜺𝟎 Initial failure strain
𝜺𝒇 Final failure strain
𝜺𝒊𝒊 Strains in the longitudinal or transverse directions (𝑖 = 1,2)
𝜺𝟎𝒊𝒊 Initial failure strains in the longitudinal or transverse directions (𝑖 = 1,2)

𝜺𝒊𝒊
𝒇 Final failure strains in the longitudinal or transverse directions (𝑖 = 1,2)

𝜺𝒊𝒊
𝒇𝒕 Tensile final failure strain in the longitudinal or transverse directions (𝑖 =
1,2)
𝜺𝒊𝒊
𝒇𝒄 Compressive final failure strain in the longitudinal or transverse directions
(𝑖 = 1,2)
𝒍𝒄 Characteristic length of the element
 Displacement
𝜹𝒐 Displacement of damage initiation
𝜹𝒇 Displacement of element at failure
 Nominal volumetric compressive strain
𝜼𝑩𝑲 Benzeggagh-Kenane coefficient mode-mix exponent
𝜞 The Grüneisen ratio
𝜞𝟎 A material constant to define the Grüneisen ratio
𝝆 Density of the gelatine projectile
𝝆𝟎 Reference density of the gelatine projectile

XX
Abbreviation

3-D Three-dimensional
APC Aromatic Polymer Composite
ASTM American Society for Testing and Materials
B-K Benzeggagh-Kenane
CAI Compression-After-Impact
CDM Computational Damage Mechanics
CF/Epoxy Carbon Fibre Reinforced Epoxy
CF/PEEK Carbon Fibre Reinforced Poly (ether-ether-ketone)
CFRP Carbon Fibre Reinforced Polymer
CT Computed Tomography
DA Damage Area
DIC Digital Image Correlation
EPFM Elastic-Plastic Fracture Mechanics
FAA Federal Aviation Administration
FE Finite Element
FOD Foreign-object Damage
FRP Fibre Reinforced Polymer
FVF Fibre Volume Fraction
HDPE High-Density Poly-ethylene
IE Impact Energy
ISO International Organization for Standardization
KEA Kinetic Energy Absorption
NDT Non-destructive Testing
OOP Out-of-plane
PEEK Poly-ether-ether-ketone

XXI
ABBREVIATION

PPS Poly-phenylene-sulphide
PTFE Poly-tetra-fluoro-ethylene
SPH Smoothed Particle Hydrodynamics
UD Unidirectional
VCC Virtual Crack Closure

XXII
Chapter 1. Introduction

1.1 Background
Since carbon fibre and Kevlar fibre were developed in the last few decades, composite
materials have played an increasingly important role throughout industrial development. The
usage of composite materials has permeated into daily lives and been expected to contribute to
a tremendous global cross-sector market of $105.8 billion in 2020 [1]. The outstanding
performances, such as high strength-to-weight ratio, no corrosion, excellent fatigue resistance,
temperature tolerant to excessive heat, low thermal expansion [2–7], of composite materials
prompted such materials as competitive alternatives to the conventional materials, i.e. metallic
materials. Nowadays, composite materials are critical to many industries and have been
extensively used in numerous scientific and commercial applications, especially in
transportation industry, where the weight of the structure has direct influence on operating cost
and performance. Example applications in railway, automotive and aerospace industries are
shown in Figure 1-1 [8–10].

(a) (b) (c)

Figure 1-1. Applications of composite materials: (a) Bombardier Transportation C20


metro train with composite car body; (b) a Formula One car with carbon fibre chassis;
(c) Airbus A350-XWB aircraft employed 53 wt% of composite materials [8–10].

1
CHAPTER 1

In aerospace industry, as shown in Figure 1-2 [10,11], the use of composite materials
on aircrafts has steadily grown since Airbus applied such materials on the secondary structures
of the world’s first twin-engine widebody aircraft A300. Until now composite materials have
been utilised extensively within primary structures of aircraft such as outer wing and fuselage.
After nearly a century of dominance as the material of choice for aircraft, metals are
seeing increased competition from composite materials in use on such aircraft as Airbus A350
XWB and Boeing Dreamliner 787 jetliner, of whose composite materials account for 53 wt.%
[10] and 50 wt.% [11], respectively. Thanks to the use of composite materials on such aircrafts,
it is not only improving flight performances but also reducing global environment impact.

(a) (b)
Figure 1-2. Composites in aerospace industry: (a) composite structural weight development on
Airbus aircrafts; (b) composite solutions applied throughout the Boeing Dreamliner 787 [10,11].

The great performance of composite materials supports the revolution of commercial


aircraft. Correspondingly, the need from aerospace industry has significantly drove various
types of polymer matrix to be invented and developed. In broad generalities, thermoset matrix
composites are more predominant than thermoplastic composites for a long history, because
they frequently have lower raw material costs and are easy to be processed and formed into
complex geometries. However, in recent years, thermoplastic composites have an increasing
market share in the aerospace industry, due to their excellent properties of higher strength and
impact resistance, better recyclability and eco-friendly manufacturing [5–7]. In other word,
they offer aircraft designers an opportunity to utilise less material to satisfy the structural
strength requirement so that the overall weight of an aircraft can be further reduced.

2
CHAPTER 1

However, the damage mechanisms of these thermoplastic composite, subjected to


various damage loading, i.e. impact loading, are not fully understood yet [12–16]. Foreign-
object debris (FOD) damage, one of the model critical damage mechanisms to thermoplastic
composites, poses serious threats to aircraft, as such damage can, lead to damage which may
not be necessarily visible but which may, nonetheless, lead to a significant structural strength
reduction. Additionally, in the worst case, lead to major structural damage in critical aircraft
components, as in Figure 1-3 [17,18]. The US Federal Aviation Administration (FAA) stated
in July 2019, that there were 214,048 bird-strikes reported to the FAA from 1990 to 2018 [15].
More than two-hundred and sixty three aircraft were seriously damaged, and over two-hundred
and eight two people were killed due to wildlife strikes from 1988 to 2018 [15]. In the case of
runway debris hazards, the Boeing reported that the resulting damage to airplanes, equipment
and personnel is estimated to cost approximately four billion dollars annually [19].

(a) (b)
Figure 1-3. Aircraft structures collapsed after (a) hail impact, or (b) bird-strike [17,18].

As a result of the increasing number of these incidents, more attentions have been paid
to the impact damage resistance of these structures on aircrafts which are constantly threatened
by FOD under not only low and high velocity impact but also the soft and hard impact. Both
the damage mechanisms and structural integrity need to be studied experimentally and
numerically to obtain a better understanding of the impact behaviour of composite materials,
which will then contribute to the wider application of composite materials in the latest
generation of aircraft.

3
CHAPTER 1

1.2 Aims and objectives


In this Ph.D. project, the main aim is to seek experimental and numerical methods to
evaluate the performances and investigate the damage mechanisms of thermoplastic and
thermoset composites under impact loading, which can be finally employed in the future
industrial applications, to assist the design and maintenance of composite structures. To
achieve the objectives, the research programme has been divided into the distinct tasks, which
are detailed as below.

i. To conduct a comprehensive literature review and learn about the state-of- the-art
in impact testing and modelling on polymer composites.

ii. To design and manufacture thermoplastic poly(ether-ether ketone) and thermoset


epoxy composite specimens to study mechanical behaviours under impact.

iii. To investigate the impact performance of unidirectional and woven-fabric


composite specimens under low velocity (i.e. <10 m.s-1) and high velocity (i.e. up
to 100 m.s-1) impact.

iv. To study the perforation behaviour of such composites impacted at high-velocity


using soft and hard impactors.

v. To develop a high-fidelity computational model to predict the impact behaviour of


composite structures under such impact.

vi. To quantitatively analyse the loading responses and failure mechanisms of the
investigated thermoplastic and thermoset composites based on the results obtained
from the impact experiments and simulations.

1.3 Structure of the thesis


The chapters in this thesis are outlined as follows.

Chapter 1 introduces the background, the objectives and the structure of this Ph.D.
project.

Chapter 2 addresses the fundamentals of fibre-reinforced composite materials,


including classification, fibre reinforcements, polymer matrices, manufacturing methods and
typical failure modes, as well as the advanced impact performance assessment methods,

4
CHAPTER 1

including characterisation testing methods, non-destructive testing (NDT) methods and


numerical modelling of fibre composites.

Chapter 3 introduces the unidirectional and woven-fabric carbon fibre composites


employed that are based upon a thermoplastic poly(ether-ether ketone) matrix (termed
CF/PEEK) or a thermoset epoxy matrix (termed CF/Epoxy), detailing their mechanical
properties and manufacturing processes, and the experimental equipements and technologies,
including drop-weight tower, gas-gun, three-dimensional (3D) Digital Image Correlation (DIC)
and ultrasonic C-scanning.

Chapter 4 details the low velocity and high velocity impact experiemnts that are
performed using a rigid impactor, at different enegy levels, using unidirectional thermoplastic
and toughened-thermoset composites which possess the same cross-ply lay-up. All impacted
specimens are inspected using ultrasonic scan device to evaluate the impact damage. The
measured loading response, energy dissipation and damage area are employed to compare the
impact performance of these tested thermoplastic and thermoset composite specimens.

Chapter 5 presents an elastic, two-dimensional finite-element (FE) model, which is


computationally efficient, for simulating both the low velocity and high velocity impact tests
on the thermoplastic and thermoset composites in Chapter 4. The modelling results are
quantitatively validated against the experimentally-measured impact responses of the tested
composites.

Chapter 6 investigates the performance of woven-fabric carbon fibre composites under


high-velocity impact loading at impact velocities of up to 100 m.s-1. The composites are
impacted using soft and hard projectiles. Both visual inspection and C-scanning examination
are conducted to assess the extent of impact damage in the tested specimens. The perforation
resistance and energy absorbing capability of the woven thermoplastic and thermoset
composites are studied by performing high-velocity impact experiments using the hard-
projectile and the resulting modes and extensts of damage are identified.

Chapter 7 presents experimental and numerical studies on the behaviour of woven-


fabric composite laminates subject to impact loading by soft gelatine projectiles. High velocity
impact experiments are performed to investigate the effect of the matrix system on the impact
response of the composites. A FE model is developed to simulate the impact by a soft projectile
on the composite specimens. Good agreement is shown between the predictions from using the
FE model and the experimental results.

5
CHAPTER 1

Chapter 8 summaises the main outcomes of this Ph.D project.

Chapter 9 suggests the future work to continue this study.

6
Chapter 2. Literature review

2.1 Introduction
In this chapter, the fundamentals of composite materials, including classification, fibre
reinforcement, polymer matrix, fabrication methods, as well as an overview of the past and
current research work on the composite materials. This overview addresses the impact
performance of the composite materials under low-velocity and high-velocity impact loading
and introduces the typical failure modes shown by composite materials. This is followed by
presenting the Non-Destructive Testing (NDT) methods for assessing the structural damage
occurred in composites. The numerical approaches for high-fidelity modelling of damage in
composites, for enabling the predicting of impact behaviour, are discussed at the last part.

2.2 Composite materials


2.2.1 Classification of composite materials
Composite materials are composed of two or more constituent materials with
significantly different advantageous properties that, when combined, produce a material with
characteristics different from the individual components. Among the constituent materials, one
is called reinforcement which provides strength and stiffness of composite, and the one in
which it is embedded is called matrix which gives overall shape and contribute to the toughness
of composite [4,7]. Such constituents remain their identites and possess an physically identical
interface between each other in the composite. A typical structure of composite is illustrated in
Figure 2-1.

7
CHAPTER 2

Figure 2-1. Typical structure of composite [4].

Composites are generally classified by either the geometry of the reinforcement or the
type of matrix. Particulate composites, flake composites and fibre composites are commonly
classified by the geometry of the reinforcement and designed for variuos application purposes
[20]. Among these composites, due to the high aspect ratio between length and diameter of
fibres, which can provide effective shear stress transfer between the matrix and the fibres, fibre
reinforced composites are most widely used in the industrial applications and are emphasised
in this thesis. Such composites can be further classified by the type of matrix, e.g. polymer,
metal and ceramic [21]. Considering the excellent properties of high strength and friendly
manufacturing principles, polymer matrix composites are the most common advanced
composites in industries [21].

2.2.2 Fibre reinforcement


2.2.2.1 Types of fibres
In a composite structure, fibre reinforcement gives strength, stiffness and other
mechanical properties to the composite, and dominate other properties, e.g. the coefficient of
thermal expansion, conductivity, and thermal transport. In the history of fibres, natural fibres
were the first uses of fibre reinforments when Egyptians and Mesopotamians combined straw
and mud for building construction. Along with the arrival of modern era, plastics were
developed and further improved by Slayter Games to enhance the strengh and rigidity of
composite materials with glass fibres [22]. Since carbon fibre (also known as graphite fibre)
and aramid fibre (known as Kevlar) were developed in 1970s [23,24], Fibre Reinforced
Polymer (FRP) became the indispensable part of composite material system. In present time,
various types of carbon fibre, glass fibre and aramid fibre are most commonly used fibres which
have different properties, for example, as shown in Figure 2-2.

8
CHAPTER 2

Figure 2-2. The diagram of tensile stress-strain for various reinforcing fibres [3].

Among these fibres, glass fibre is the most common fibre applied polymer matrix
composites. The advantages of glass fibre include low raw material cost, relatively high
strength, high toughness, excellent heat and corrosion resistance [25]. In contrast to glass fibre,
carbon fibre exhibits great improved strength and stiffness at a lower weight, but with a higher
material cost. Such fibre contain at least 90 wt.% of carbon and is produced from polymeric
presursor materials. From a macro perspective, carbon fibre is very thin filament with a diamtre
of approximately 5-10 μm, which can transfer shear stress effectively between the matrix and
fibres. Furthermore, carbon fibre provides excellent resistance to corrosion, chemical and
fatigue. Such properties make it an outstanding choice as fibre reinformcent for the advanced
load-bearing applications [26]. Similarly, Kevlar offers high specific tensile strength, excellent
abrasion and fatigue resistance. However, it has poor compressive strength and degraded
peroformance against ultraviolet radiation or acid corrosion.

9
CHAPTER 2

2.2.2.2 Fabric Architecture


Fabric architecture is aimed as the structural arrangement of fibres in a composite,
which not only influences the properties of the composite, but also its processing. In general,
the fibre layers in various forms can be embedded into a matrix either in continuous lengths or
discontinuous lengths. Continuous FRPs normally have a preferred orientation and high aspect
ratios, while discontinuous fibres have a random orientation and low aspect ratios, which
dramatically reduces their strength and modulus [27]. Continuous FRP composite is often made
of multi layers of the fibre-reinforced plastics by stacking the unidirectional (UD) fibre or
woven fibre layers and curing them into the desired thickness at certain temperature and
pressure. Because of this, continuous FRP composite provides good strength in axial loading
but less in transverse loading [25].

UD fibre composites are the most commonly used cotinuous FRPs in commercial
aerospace due to their exceptional specific properties. The in-plane performance of UD
composites is generally superior to any other class of FRP, e.g. woven-fabric. The reason for
this is largely attributable to the way fibres are arranged within these materials. A lamina is the
building block of FRP structures, which may be oriented in different directions that will
enhance the strength in the primary load direction. Figure 2-3 illustrates how fibres in UD
laminates might be arranged. UD prepregs have highly aligned fibres with a regular packing
arrangement that allows for a high Fibre Volume Fraction (FVF). These properties give UD
laminates their exceptional in-plane properties. In contrast, a woven-fabric lamina is usually
offered as a cross-ply construction. Weaves are made on a loom by interlacing two orthogonal
sets of yearns. This specific waviness of woven-fabric composites attributes to the most
resistant to in-plane shear movement, but reduces the strength and stiffness of the composite
[27]. Note that one major problem with multilayered laminates is that their interlaminar
properties can be low and they can be prone to early failure by delamination, in which cracks
originated at the interface between the layers due to high interlaminar tensile and shear stresses
cause separation of layers [26]. To overcome this problem, 3D weaving structures have been
developed to provide out-of-plane reinforcements and produce composites with enhanced
through thickness properties. However, this structure may cause negative influence on the in-
plane fibres and thus the degraded primary properties of composite.

10
CHAPTER 2

Figure 2-3. Illustration of fibre alignment in two layers of a cross-ply UD prepreg laminate
[25].

2.2.3 Polymer matrix


Matrix, in general, gives shape to the composite part, protests the reinforcements from
the environment, transfers loads to the reinforcements, and contributes to mechanical properties,
e.g. toughness, that depend upon both the matrix and the reinforcement [3]. Polymer matrix, as
the most popular matrix employed in the advanced high performance FRP composites, can be
classified into two categories: thermoplastics and thermosets, according to their mocular
structures which are illustrated in Figure 2-4.

Figure 2-4. Molecular structure of thermoplastic and thermoset polymers [28].

11
CHAPTER 2

Polymers are composed of long-chain molecules consisting of many repeating units. In


thermoplastics, linear polymer arrays with amorphous, semicrystalline or mixed morphology
exhibit weak intermolecular forces between polymer chains. No chemical bonding or cross
links usually take place during the fabrication process. Therefore, thermoplastic resins can be
repeatedly moulded when exposed to heating [2,5]. In thermosetting resins, strong covalant
bonds from cross links between polymer chains are formed from an irreversible chemical
reaction during the fabrication process. The cured or crosslinked material cannot be
reprocessed even heated [3].

Due to the crosslinked polymer networks, thermosets exhibit excellent heat and
chemical resistance, but with a resultant compromise of characteristics including formability
and toughness. They are relatively brittle, resulting in low fracture toughness, reduced damage
tolerance and low transverse tensile, interlaminar tensile and in-plane shear failure strains.
Thermosets are also prone to microcracking during processing and when subjected to
mechanical loading[26]. On the other hand, thermosets offer advantages over thermoplastics
in east of processing, addition polymerisation with no evolution of volatiles, high creep
resistance, good fluids resistance and low resin costs.

In contrast, thermoplastics typically offer excellent toughness, high damage tolerance


and excellent shelf life. However, their resistance to chemical, heat and creep are lower than
thermosets[27]. In addition to thermoplastics, high molecular weights common in high-
temperature thermoplastics result in high melt viscosity and low flow in the matrix that can
make processing difficult. Poor management of volatiles generated from solvent or polymer
condensation reactions can also produce voids in the composite, degrading the mechanical
properties of the material. The comparison of properties between thermoplastics and thermoset
are summarised in Table 2-1.

12
CHAPTER 2

Table 2-1. The comparison of properties between thermoplastics and thermosets [20,21,29].
Thermoplastics Thermosets
Soften on heating and pressure Decompose on heating
Higher fabrication temperature and viscosities Lower fabrication temperature
Can be remoulded and recycled Cannot be remoulded or recycled
High strains to failure Low strains to failure
Short cure cycles Long cure cycles
Fair heat resistance Excellent heat resistance
High resin cost Low resin cost
Indefinite shelf life at ambient temperature Definite shelf life at low temperature

2.3 Fabrication processes and their effects


The fabrication stages for FRPs, including lay-up, moulding and curing, are covered in
this section. In lay-ups, wet lay-up, prepreg lay-up and spray lay-up are the major processes.
In the stages of moulding and curing, a number of fabrication processes [6,27], such as vacuum
bag forming, injection moulding, compression moulding and autoclave curing, have been
established for both thermoplastic and thermoset composites. Over the past few decades, the
manufacturing processes have been developed which make use of the economic processing
characteristics. In particular, the effect of fabrication process of thermoplastic and thermoset
composites on mechanical behaviour of the composites have been well studied and will be
discussed in this section.

2.3.1 Fabrication processes


For thermosets, prepreg lay-up, vacumm bag forming and autoclave curing are the
predominant fabrication method used in the aerospace inductry. Typically, a prepreg tape, fibre
reinforcement impregnated with a partically-cured thermosetting resin, is stored at low
tempearture around -18 °C as the viscosity and tack of the thermosetting resin is quite sensitive
to temperature and humidity. After prepreg lay-up, the lamiante is sealed in a vacuum bag for
curing, as shown in Figure 2-5. Vacuum bag forming and then cured at an autoclave. The
aerospace industry mostly uses carbon-epoxy prepreg which is designed to be cured at around
175 °C to 180 °C.

13
CHAPTER 2

Figure 2-5. Schematic of a vacuum bag system [4].

In overall, the fabrication of thermoplastic composites is more friendly to be handled.


Unlike thermosets, which undergo a chemical curing reaction for geometry stabilisation during
forming, the forming process for thermoplastic composites is relatively simple and involves
purely physical changes to the material. Fully consolidated prepreg tapes are mostly used and
processed via compression moulding, as shown in Figure 2-6. The compression moulding is
used here to denote processes where the forces required to form a part are provided by a high-
pressure press, containing matched metal plates. The compression moulding process is non-
isothermal and capable of high production rates with cycle times typically under ten minutes
but at elevated temperature between 280 °C to 400 °C.

14
CHAPTER 2

Figure 2-6. Schematic of a compression moulding system [3].

2.3.2 Effects of manufacturing process


During processing, matrix flow through the fiber architecture determines the void
content, fibre wetting, fibre distribution, dry area and others in the final composite, which in
turn, also affect its properties and performance. The industries therefore have increasingly
focusing on optimising the composites properties and improving the ecomonical efficiency.

Thermoplastic composites due to their inherent properties, a number of fabrication


processes have provided a cost-effective way to enable a significant reduction in manufacturing
cycle times by eliminating the time-consuming curing step which is necessary for thermosets.
The influence of the consolidation temperature, holding time, contact pressure, cooling rate on
the mechanical properties of the thermoplastic composites have been investigated and the
optimum mechanical properties of the composites were studied [30–39]. For example,
crystallinity, as a function of cooling rate, can lead to enhanced stiffness and promoted good
mechanical properties. The crystallisation of PEEK depends on first on the state of the melt
from which it is crystallised. It is preferred that this melt should be at a temperature at least
360 °C [31]. The crystallisation of the matrix slowed down after time in the melt state but the
crystallisation behaviour cannot be affected by longer period heating, which concluded by
Saliba et al. [32] who held times of 2 hours at 400 °C for Aromatic Polymer Composite (APC)-
2. Leicy et al. [33] have studied the impact properties of standard processed and quenched

15
CHAPTER 2

mouldings of carbon fibre/PEEK. They observe that the incompletely crystallised samples
absorb 10% less energy at punch through, but that they are significantly less prone to
delamination on impact. Fujihara et al. [34] have shown that a relative lower consolidation
temperature (380-410 °C) with a short holding time (20 min) will achieve better mechanical
performance for APC-2 prepregs. Moreover, Jar et al. [35] have revealed that the interfacial
strength between matrix and fibre will be improved while the consolidation temperature
increased from 380 °C to 400 °C, but this property will decrease dramatically if the temperature
increased over 400 °C.

For thermoset composites, it is well known that the fabrication of thermoset composites
is mainly depend on the autoclave process, which requires long cycle time and high energy
consumption. Therefore, many investigators [36–39] have been focused on the effect of cure
cycle time and temperature in autoclave processing. Kim and Daniel [37] studied the effect of
curing cycle on the development of material properties and claimed that the residual stress
developed at the excessive curing temperature due to the constraint-induced strain. Coenen et
al. [38] developed a novel out-of-autoclave process for CF/epoxy composites, and Davies et al.
[39] tested this process and claimed that the flexural strength of the novel processed specimen
was found to be 10 % lower than the specimen manufactured by conventional autoclave process.

2.4 Impact performance


Impact performance of composites materials is deemed as a critical and fundamental
field in the aerospace industry. The impact strength of a composite material is the ability of the
material, under impact loading, to absorb mechanical energy without considerable deformation
occurring [40]. Before considering the effects of impact loading on composite materials, it is
wise to consider the different forms of impact loading and typical failure modes that can be
achieved. Impact loading can be classified into low and high velocity impacts, whereby
researchers [41–44] have collectively agreed that the threshold velocity between the two forms
of impact loading, which also defines the threshold of low velocity impact, is 10 m.s-1. Davies
and Robinson [45] provide a more detailed definition of low and high velocity impacts,
whereby they define a low velocity impact as one in which through-thickness stress waves in
the specimen play no significant part in the stress distribution at any time during the impact
event, whereas a high velocity impact is stress wave dominated. Another classification was
based on the structural response of the impacted plate. Thereby, low velocity impact speeds

16
CHAPTER 2

were impact speeds at which the contact duration between impactor and target was long enough
for the structure to deform in phase with the impactor [46], as illustrated in Figure 2-7. In
contrast to that were high velocity or ballistic impacts where material damage was very
localised and mainly caused by through-thickness stress waves.

Figure 2-7. Schematic representation of impact response under low and high-velocity
impact [46].

Low velocity impact is usually classified as being due to dropped tools used during
maintenance services, whereas hailstorms, runway debris impacting the fuselage during take-
off and landing, bird strikes and ice from propellers striking the fuselage are sources of high
velocity impact [44]. Various modes of damage can occur due to the sources of impact
mentioned above.

2.4.1 Under low-velocity impact loading


The low-velocity tests are generally performed using a drop weight or a swinging
pendulum. A number of standard testing procedures, e.g. International Organization for
Standardization (ISO), American Society for Testing and Materials (ASTM) standards, are
available to assess the impact performance of the composite materials by applying the standard
on specimens of fixed size and geometry, and testing conditions employed. There are many
data, such as load-time curve, load-displacement curve, can be obtained from a low-velocity
impact test via the data acquisition system equipped.

17
CHAPTER 2

Figure 2-8. Schematic of drop-weight impact fixture in the ASTM 7136 standard [47].

2.4.1.1 Effect of matrix resin on impact performance

Considering previous research on the impact resistance of thermoplasticand thermoset


composites, as early as 1985, researchers [48–50] first reported that thermoplastic composites
possessed a superior resistance to impact damage, when subjected to relatively low-velocity
drop-weight tests, compared with the then currently available CFRPs which employed a matrix
of a relatively very brittle, thermosetting epoxy polymer. Vieille et al. [51] have undertaken an
experimental study to compare the performances of a CF/Epoxy composite and carbon-fibre
reinforced poly(phenylene sulphide) (PPS) and CF/PEEK composites, all with a woven-fibre
architecture, when subjected to a relatively low-velocity impact using a drop-weight test. They
found that the CF/Epoxy composite experienced a larger area of delamination damage
compared with the thermoplastic-matrix CF/PPS and CF/PEEK composites. They suggested
that this reflected the tougher nature of the thermoplastic PPS and PEEK matrices compared
with the thermosetting epoxy matrix. This also further confirmed that a tougher matrix system
can facilitate composite laminates to present a better impact performance.

Moreover, Chamis et al. [52] executed a range of Izod tests on composite structures.
They concluded that ductile thermoplastic composites are considered tougher than thermoset
composites, that are considered brittle due to their brittle thermoset matrix such as epoxy resin,
due to thermoplastic composites exhibiting a greater area underneath their stress-strain curve
when compared to thermoset composites. This agrees with recent observations made by
Sundaram et al. [53], which states that thermoplastic composites, such as CF/PEEK, who fail

18
CHAPTER 2

in a ductile manner and hence possess a large area under their stress-strain curve, possess
greater mechanically stability during impact and greater fracture toughness than thermoset
composites. This is due to the plastic nature of the thermoplastic composite’s resin which
causes the impact damage to propagate in a ductile manner through the composite. This
statement is further supported by Jogur et al. [44] and Faur-Csukat [54], whereby they both
managed to conclude that thermoplastic composites possess a greater impact resistance than
brittle thermoset composites, due to their higher toughness characteristics.

2.4.1.2 Effect of laminate lay-up configuration and projectile geometry


on impact performance

Caminero et al. (52), who investigated the impact performance of unidirectional and
multi-directional laminate stacking sequences, found out that a sequence of laminate plies
oriented at ±45° exhibited the best impact performance out of all the orientations tested, due to
the fact that the impact load is shared between the matrix and fibre reinforcement. This finding
was reinforced by Hu et al. (58), who stated that as the proportion of ±45° layers of fibres
increase, the energy absorption of the laminate increases, while the maximum deformation
exhibited by the composite structure decreases.

Mitrevski et al. [17] have reported the effect of impactor shape on the impact behaviour
of woven CF/Epoxy composites. Their studies involved the low-velocity impact of the
composites using 12 mm diameter hemispherical, ogival and conical head-shaped impactors at
impact energies of 4 J and 6 J. They found that the energy absorbed by the specimen was the
highest for the conical impactor which also produced the largest penetration depth. However,
the peak load was greatest for the hemispherical impactor which also produced the shortest
contact duration. In a subsequent paper [18] it was reported that different impactor head-shapes
produced different levels ofthe various types of damage that were observed, such as fibre
breakage, matrix cracking and delamination, which affected the residual properties of the
composite. It was found that the relatively blunt hemispherical impactor produced the largest
damage area and the type of damage was dominated by delamination. Robinson and Davies
[19] have reported results from drop-weight tests using a hemispherically-shaped impactor at
impact velocities ofup 6.0 m.s-1 striking woven glass- and carbon-fibre composites based on a
matrix of an acrylic polymer. They found that the impact damage caused was mainly a function
of the impact energy, and not the velocity or mass, separately, of the impactor. Further, the

19
CHAPTER 2

measured peak load was also shown to correlate with the impact energy independent of the
impactor mass used, indicating that the low-velocity impact of these specimens was a quasi-
static process. Their test results also indicated that the peak load versus damage size
relationship was independent of the diameter of the impactor. More recently, Sevkat et al. [20]
have undertaken drop-weight tests at an impact velocity of 4.1 m.s-1 on woven glass-
fibre/carbon-fibre hybrid epoxy composites. They reported that the contact area between the
impactor and the composite test specimen was a critical factor in the response ofthe composites.
Impactors with relatively large contact surfaces produced a higher initial load peak, a higher
maximum load, a shorter contact duration for the time-scale of the impact event and more
extensive delaminations between the hybrid layers.

2.4.2 Under high-velocity impact loading


In contrast to the low-velocity test, at present, no acceptable standard testing procedures
are available for high-velocity impact testing of composite materials. However, over the past
few decades, numerous research has been undertaken in an attempt to better understand the
impact response and failure mechnisms of CFRPs under high-velocity impact. As Cantwell and
Morton [41] claimed that it is difficult to separate the effects of mechanical properties, such as
strength and stiffness of the composite matrix and reinforcement respectively, and geometrical
factors, such as projectile shape and laminate lay-up configuration, on the overall impact
performance of the composite structure. However, it is very beneficial to present various
observations made by numerous researchers on the sole effect that mechanical properties of the
composite, the projectile geometry and the laminate lay-up configuration have on the impact
performance of the composite material.

2.4.2.1 High-velocity impact responses

The physical phenomena occurring in laminates under high velocity impact is deemed
complex. As the effect of velocity or intensity of dynamic loading is increased, the potential
for damage and subsequent reduction in mechanical performance is increased. Numerous
observations were made by various researchers that conducted high velocity impact simulations
and experiments on composites. In terms of high strain rate effects, a number of researchers
observed that the tensile and compression modulus and strength increased with the increase in

20
CHAPTER 2

the strain rates. However, the shear modulus and strength were discovered to decrease with the
increasing strain rates, and the strain rate effects in the composite appeared to be influenced
primarily by the matrix viscoelasticity, fiber-matrix interfacial properties, the composite woven
reinforcement architecture, and the time-dependent nature of damage accumulation. Jogur et
al. [44] suggested that the dominating factor in high velocity impact is the propagation of stress
waves throughout the target during the impact event. The duration of impact during a high
velocity impact event between the impactor and target is very low, and hence does not allow
stress waves to propagate through the thickness of the laminate and reach the rear surface. This
idea of the stress waves not reaching the rear laminate surface was challenged by Will et al.
[55], as they revealed that the stress waves reach the rear laminate surface and get reflected
forming tensile waves, which act on neighbouring ply interfaces initiating delamination.
Additionally, they concluded that a larger damage area is found on the rear surface, due to
delaminated layers sliding over each other as the impact projectile reaches them, thus causing
the layers of laminate to fail completely due to tensile failure. This agrees with Abrate [56]
damage area depiction in Figure 2-9 which is further supported by Raimondo et al. [57]. This
idea was supported by Cantwell and Morton [46] and Higuchi et al. [58], who both discovered
that stress wave reflections from the rear laminate surface caused a phenomenon called spalling,
which involves the removal of a large number of the outermost fibres.

Figure 2-9. Illustration of impact damage in a laminate [55].

2.4.2.2 Effect of matrix resin on impact performance

Many reasearchers have investigated the effect of matrix system on impact performance
of composites. Among the composites, thermoplastics are emerging in the aerospace industry

21
CHAPTER 2

as promising competitors to typical thermosets thanks to their higher impact resistance and
toughness. Morita [59] investigated the impact resistance of CF/PEEK and CF/epoxy
composites when subjected to a high-velocity impact using a hard-projectile with a
hemispherical end, and claimed that the CF/PEEK exhibited superior impact resistance than
the CF/Epoxy. Wagner et al. [60] performed similar high-velocity impact tests and concluded
that, at high impact energy level, CF/Epoxy laminates exhibited similar size of damage area to
the CF/PEEK speicmens but with servere delamination behaviour. This may lead to a reduction
of the structural strength of the specimen. However, several researchers challenged the fact that
the impact resistance of composites is reliant on the mechanical properties of the matrix.
Husman et al. [61] discovered, after executing high velocity impact tests on carbon fibre
composites, that the impact performance of the composite material does not depend on the
toughness of the polymeric matrix. Moreover, Cantwell [41] investigated the independence of
the impact performance on the toughness of the polymeric matrix. He suggested that, in brittle
composites, the resin toughness is fully transferred to the composite, however, in ductile
composites, the resin toughness is not fully transferred to the composite and therefore the
positive effect of having a tough matrix is reduced. This idea has chanllenged by various recent
studies that suggest methods by which the mechanical properties of composite structures can
be enhanced [62–65]. Bull et al. [65] suggested that the toughness of composites with
thermosetting resins can be improved by adding a toughener, causing the composite to have
three phases, consisting of the thermosetting matrix, the fibre reinforcement and the toughener.
This was supported by Nash et al. [62], who confirmed that by modifying a thermosetting
composite through the inclusion of a thermoplastic phase, the impact performance of the
composite can be improved and enhanced.

2.4.2.3 Effect of laminate lay-up configuration on impact performance

A number of studies have investigated the effect of the laminate lay-up configuration
on the impact performance of composite structures [55,60,66,67]. For example, Cantwell and
Morton [66]studied the effect of laminate stacking sequence on the impact performance of
CFRP composites. They determined that the extent of delamination observed in the composite
material was related to the energy absorbed during impact and was affected by the orientation
angles of laminates stacked on top of each other. This was explained and supported by Will et
al. [55], who suggested that delamination is more likely to occur at interfaces where there is a

22
CHAPTER 2

mismatch in stiffness. Since a mismatch in stiffness occurs at the interface of fibres oriented at
different angles, delamination is deemed to be highly likely to occur at these interfaces. Wang
et al. [67] determined that as the fibre orientation angle increases, the debris from the high
velocity impact decreases in size. It was suggested that this is due to the fact that the
circumferential stiffness of the composite increases with increasing fibre orientation angle, and
hence adds resistance to interlaminar crack growth. Additionally, it was observed that the
laminates with low fibre angles were less prone to fracture than the laminates with greater
values of fibre angles.

2.4.2.4 Effect of projectile geometry on impact performance

Many researchers have investigated the effect of projectile geometry on the impact
performance of composite structures [68–70]. Appleby-Thomas et al. [68] observed, by
conducting a series of high velocity impact tests on composites structures, that blunt nosed
projectiles caused extensive deformation and fragmentation of the target composite, due to
plugging and shearing of the laminate. On the other hand, for projectiles of sharper nose
geometries, they observed that the projectile had penetrated the composite to a greater degree
in comparison with the blunter nosed projectiles. Therefore, it was concluded that as the
sharpness of the projectile nose increases, less kinetic energy is lost by the projectile, hence
allowing it to penetrate the laminate more efficiently, causing a smaller damage area in
comparison with blunter nosed projectiles. Conversely, Babu et al. [69] conducted a series of
high velocity impact tests using hemispherical, conical, and conical with round nose and
truncated tip projectiles. It was observed that the projectile nose geometry had insignificant
influence on the damaged area on the composite target, which agrees with Seifoori’s et al. [70]
observation. Furthermore, they were able to establish from their experimental results that the
maximum amount of energy absorbed by the composite was exhibited in the target plate struck
by the truncated tip projectile. This is due to the sharper conical projectiles having a point
contact between them and the target, whereas for the truncated nose projectile, the contact area
is greater, and hence the load will be distributed across the flat portion of the tip. Consequently,
this causes higher impact energies to be needed in order to observe laminate penetration for
truncated tip projectiles in comparison with normal conical projectiles.

23
CHAPTER 2

2.4.3 Typical failure modes


The failure modes, presented by the fibre reinforced composite laminates, are generally
fibre fracture, matrix cracking and delamination [43,71], Figure 2-10. In these three types of
failure, the fibre fracture and matrix cracking can be summarised as intralaminar damage, and
the delamination can be defined as interlaminar damage [72,73]. Furthermore, all failure modes
contain two processes: initiation and propagation. Laffan et al. [74]pointed out that the critical
strain energy release rates associated with these failure modes are properties that are intrinsic
to the material system, which need to be measured for completely characterising the damage
tolerance of the material system [75–77]. Due to the properties of matrix material, composite
laminates tend to present nonlinear behaviour under shear loading, which has significant effects
on the performance of composite structures, and should also be accurately characterised [78–
80].

(a) (b)
Figure 2-10. The main failure types shown in the unidirectional FRP laminates: (a) experiemtnal
observed failure modes [43] and (b) schematically categorised failure modes.

2.4.3.1 Interlaminar failure


Delamination is a type of failure between plies and may result in significant loss of
compressive strength and stiffness of the laminate [81,82]. This is due to the fact that
delamination has the effect of dividing the laminate into sub-laminates that are more
susceptible to buckling when loaded in compression. The total strain energy release rate
associated with the mode of failure, can be decomposed into 𝐺𝐼 , 𝐺𝐼𝐼 and 𝐺𝐼𝐼𝐼 based on the
different failure modes: Opening mode (Mode I), Shearing mode (Mode II) and Tearing mode

24
CHAPTER 2

(Mode III), as shown in Figure 2-11, relating to interlaminar tension, sliding shear and tearing
shear, respectively [81,83,84].

Figure 2-11. Modes of interlaminar failure [81].

2.4.3.2 Intralaminar failure


The failure modes of intralaminar failure consist of intralaminar fibre
tensile/compressive failure and matrix tensile/compressive failure [85–87]. In a unidirectional
composite ply, the material response in the longitudinal (0˚) direction is dominated by the fibres.
In woven plies, the fibres dominate the material response along warp (0˚) and weft (90˚)
directions. For fibre-dominated failure, the large amounts of fibre-matrix de-bonding and
subsequent fibre pull-out contribute to the homogenised ply-level fracture toughness [88,89].
For fibres in compression, failure may initiate from small misalignments to the loading axis
which propagate as shear driven fibre failure and kinking.

(a) (b)
Figure 2-12. Fibre-dominated (a) tensile and (b) compressive failure modes [90].

25
CHAPTER 2

For matrix-dominated failure, fracture plane will be formed on the specimen under
tensile or compressive loading perpendicular to the fibre direction. As exampled in Figure 2-
13, the failure patterns of the woven-carbon/epoxy specimens tested under tensile and
compressiove loading in the through-thickness direction [91,92]. Both tensile and compressive
failure showed fracture planes perpendicular to the loading direction.

(a) (b)
Figure 2-13. Fracture planes obtained from woven CFRPs under (a) tensile and (b)
compressive loading [91].

2.4.3.3 Nonlinear shear behaviour


Under shear loading, both unidirectional and woven carbon-fibre reinforced composites
behave non-linearly [78,93,94], which can be demonstrated by the stress-strain curves, for
example, in Figure 2-14. This nonlinear response, which is mainly caused by the yielding of
the resin, can affect the distribution of stress and the onset of failure [95–97]. The
corresponding strain localisation can be readily observed along the ligament region, where
failure modes including matrix cracking, fibre/matrix debonding, fibre pull-out and fibre
breakage can be found [94,98,99].

26
CHAPTER 2

Figure 2-14. Shear stress-strain curve for the static [+45°, -45°]2s composite specimens under
tensile tests [97].

2.5 Techniques for detecting impact damage in fibre


reinforced polymer composites
There are various defects and damage modes that occur in composite structures which
may have been introduced from the rwa materials, manufacturing, assembly or during service
due to impact loading. Damage to composite structures is not always visible under visual
inspection. Therefore, it is critical to be able to effectively detect these flaws and modes of
damage without structurally impairing the composite structures. This can be achieved using
non-destructive testing (NDT) techniques, which are an essential part of the quality assurance
philosophy of large commercial aerospace companies [100]. A number of NDT techniques,
such as ultrasonic C-scanning and X-ray Computed Tomography (CT) scanning, whereby each
possesses a set of advantages and disadvantages when compared with each other.

2.5.1 Ultrasonic C-scanning


Ultrasonic C-scanning technique has been used as one of the most widespread NDT
techniques for composite damage evaluation, and as such is often used to benchmark other
techniques [100,101]. Ultrasonic C-scanning involves transmitting sound waves through a

27
CHAPTER 2

medium which is to be tested, and then analysing the reflected sound waves. Ultrasonic waves
with frequencies ranging between 0.5 to 50 MHz are transmitted through the composite that is
to be tested, and upon encountering a discontinuity or delamination in the composite material,
the waves are reflected back and received by a transducer. Ultrasonic C-scanning results are
usually viewed in the form of either an A-, B-, or C-scan, whereby the C-scan, which is most
frequently used, is a planar view of the A scan’s displacement (of the signal through the
thickness of the material) versus amplitude information, but represented in the form of a colour
map, whereby variations in the colour map indicate the presence of flaws or defects. The A-
scan and C-scan representations of ultrasonic data are demonstrated in Figure 2-15.

(a)

(b)

Figure 2-15. Representations of (a) A-scan data and (b) C-scan data [21].

The C-scan maps, presented in the form of a colour map, provide an integrated view
and detailed depth distribution of the damage area. The position and size of a delamination can

28
CHAPTER 2

be determined by comparing the total travel time (delamination position) and amplitude
(delamination size) of the wave reflected from both the delamination and the back wall of the
material [100]. Because of this principle, the C-scan maps can only present the superposed
delamination areas in various plies. In addition to this, due to the propagating mechanisms of
impact stresses, the interlaminar fracture extends considerably from the point of impact until
the bottom ply, which causes a conical fashion from the impact point, i.e. the smaller
delamination area is in the upper (near-impact-surface) ply while the largest delamination area
is in the lowest ply [3,102]. Therefore, it is considerably important to scan from the impact
surface which would allow to determine the shape and depth of the delamination in various
plies. Moreover, the presence of delamination can be detected by observing a sudden reduction
in amplitude of the back-wall echo, while for smaller voids and inclusions, reflected signals
which are weaker than that of a crack or delamination are detected . In addition to this,
ultrasonic C-scanning has the disadvantages of relatively slow inspection speed and the
requirement of a coupling liquid.

2.5.2 X-ray Computed Tomography (CT) scanning


X-ray CT scanning relies on the computational reconstruction of radiographs acquired
using different angles of illumination on the test object, while the image contrast depends on
the differences in the attenuation of X-ray radiation passing through the investigated specimen
[103]. This popular method allows effective assessment and visualisation, in 3-D, of failure
modes such as fibre breakage, delamination and matrix cracks, thus preventing various
misinterpretations caused by previously generated 2-D images. In addition, X-ray tomography
allows the user to export composite laminate structures, in order to generate image-based
meshes, which can then be used to produce finite element models of the composite structure
[103]. However, X-ray tomography possesses numerous disadvantages, such as the fact that it
is highly expensive and time consuming [104]. In addition, the size of the test chamber greatly
limits the size of the sample to a few centimetres, causing the technique to be unsuitable for
aerospace applications and components. Also, X-ray tomography relies on producing a test
sample from the manufactured component and cannot be performed in situ [104], and therefore
is unpractical for aerospace components, as specimens cannot be obtained from an in-service
aircraft.

29
CHAPTER 2

2.5.3 Other techniques


Generally, in all NDT methods, the composite is evaluated by the detection of a test
signal, which is used to assess the structural health of the composite structure. However, none
of which are universally applicable. The choice depends on cost, ease of testing, types and size
of defect. Therefore different techniques are chosen to detect different types of flaws. A number
of other NDT techniques, such as eddy current, acoustic emission and shearography, have
different advantages on various types of damage detection.

The eddy current technique, which is classed as a radio frequency NDT technique,
induces current in a conductive material, according to Faraday’s law [105]. For the case of
composite materials such as CFRP, the induced currents flow, in the composite structure, along
fibres and from one fibre to another through the fibre contact points. The impedance of the
current flow through the fibres will change at a discontinuity or flaw in the composite structure,
and hence flaws in the composite structure can be detected using an impedance analyser
[105,106]. This technique is designed to detect delamination, fibre breakage, resin rich zones
and even observe the fibre orientations within the composite structure, but not to detect matrix
cracking [106].

The acoustic emission technique, whereby it relies on detecting, using an array of


piezoelectric transducers, the mechanical vibrations that stem from flaw and defects in the
composite material, such as matrix cracking, fibre-matrix debonding, delamination and fibre
pull-out. This technique relies on detecting the energy released from the medium rather than
supplying it with energy, which is the case with most other NDT techniques [107]. It is not
only able todetect the various modes impact damage but also fatigue damage. However, it
requires highly trained and skilled individuals to be able to use the information of mechanical
vibrations released from flaws, and correctly associate it with a specific defect or damage mode
in the composite material [107].

Shearography, is considered as one of the most advanced NDT tehcniques, relies on the
fact that light waves, of different path lengths, produce a fringe pattern when interference
occurs [101]. A shearing interferometer is used to laterally shear the reflected beam, causing
interference between the sheared and direct images [100]. If a delamination is present in the
composite material, under some form of loading, the location where the delamination is present
will deform differently than other locations in the material. Therefore, when compared with the
unloaded specimen reference image obtained, allowing detection of the presence of a

30
CHAPTER 2

delamiantion in the composite structure. However, it possesses the disadvantage of not being
able to detect and characterise other forms of impact damage other than delamination [107].

2.6 Composite damage modelling


Over the past few decades, a number of techniques have been proposed to develop
physically-based composite material damage models predicting the ultimate strength of
composite structures. Most are broadly based on distinguishing between interlaminar and
intralaminar damage, with the latter further accounting for fibre-dominated and matrix-
dominated failure modes.

2.6.1 Interlaminar modelling


Delamination, which may result from impact damage or manufacturing defects, can
lead to a significant reduction in the load-carrying capacity of composites structure [108–110].
An understanding of the mechanism of delamination and the capability to build an effective
failure model for this kind of failure deserve particular attention. Most evaluations of
delamination growth apply a fracture mechanics approach and evaluate energy release rates for
self-similar delamination growth [111–114]. At present, there are primarily two methods,
which are widely used, to predict interlaminar damage—the Virtual Crack Closure (VCC)
technique, or the use of cohesive zone model.

2.6.1.1 Virtual crack closure technique


According to the formulations and assumptions of Elastic-Plastic Fracture Mechanics
(EPFM), it is possible to predict crack growth with the VCC technique, proposed by Rybicki
and Kanninen [115]. Despite the valuable information, regarding the onset and the stability of
delamination, which can be acquired using the VCC technique, there are limitations. The VCC
technique may require complex moving mesh techniques to advance the crack front when the
local energy release rates reach a critical value [116]. VCC technique analysis also requires
that an initial delamination be defined, while for particular geometries and load situations, it is
not easy to confirm the locations of the delamination front [117].

31
CHAPTER 2

2.6.1.2 Cohesive elements model


The application of cohesive elements placed at the interfaces between plies can
overcome some of the shortcomings mentioned above. Cohesive elements are based on the
Dudgale-Barenblatt cohesive zone approach [118], which can be related to Griffith’s theory of
fracture when the size of the cohesive zone is negligible compared with the characteristic
dimensions of the structure being analysed [119]. A main advantage of using cohesive elements
is the capability to predict both initiation and propagation of delamination without previous
knowledge of the crack location and propagation direction, provided that cohesive elements
are defined at all ply interfaces.

As for the mixed mode cases, the quadratic delamination criterion is used to predict the
initiation of the softening process. The most extensively used criterion to predict delamination
propagation under mixed-mode loading is the power law criterion. Similar expression can be
derived from a more accurate criterion presented by Benzeggagh and Kenane (B-K Law) [120].
The development of a cohesive element allows a faithful description of delamination under
mixed mode and the variation of mode participation during the delamination process [121].

2.6.2 Intralaminar modelling


A number of researchers have proposed numerical models for capturing intralaminar
failure. The majority of current intralaminar damage models are based on the more empirical
approaches. For different types of damage in composite structures, intralaminar fibre failure
and matrix failure, these intralaminar models can predict the response reasonably well. More
importantly, these models provided a basis for developing more complicated cases including
crush and impact damage [122–124].

2.6.2.1 Modelling fibre-dominated damage


On the basis of Mohr’s and Hashin’s considerations on brittle fracture, Puck and
Shurmann [125] developed a new inter-fibre fracture criterion for different fracture modes in
fibre-reinforced plastics. Results relating to crack initiation and fracture as well as an indication
of the direction of cracks and the fracture angle can be gained from this phenomenological
model. By drawing on Dvorak’s [126] fracture mechanics analyses of cracked plies and Puck’s
action plane concept, Davila and Camanho [86] proposed a new physically-based criteria which

32
CHAPTER 2

can predict fibre failure without requiring parameters of curve-fitting. Specifically, an in-situ
effect was considered in their model. They made an assumption that only longitudinal shear
strength is subjected to in-situ effect. The result shows a significant improvement over the
Hashin criteria.

2.6.2.2 Modelling matrix-dominated damage


Criteria for determining a matrix fracture plane, in a unidirectional ply under various
damaging loads, were proposed by Puck and Schurmann [125]. The criteria suggested that for
transverse tension, this fracture plane will be normal to the load, while for transverse
compression or shear, the fracture plane would be determined by the ability of the matrix to
resist the shear loading. The stress state, at a fracture plane, is characterised by a normal stress,
𝑛 and shear stresses 𝑛𝑡 and 𝑛1 , as shown in Figure 2-16. A stress-based damage criterion
can be employed to capture the matrix damage initiation in composite laminates. It is generally
assumed that the fracture toughness associated with intralaminar matrix propagation is similar
to interlaminar Mode II fracture toughness [127].

Figure 2-16. Stresses and fracture plane on a composite material [125].

2.6.3 Modelling impact behaviour


A selection of FE based composite damage models is available in commercial packages.
Examples include the Abaqus built-in progressive composite damage model based on the work

33
CHAPTER 2

by Matzenmiller et al. [128]; and the LS-DYNA [129] material model type 262 which uses an
approach based on the failure criteria presented by Chang and Chang [130]. Despite the
widespread application of these commercial packages, calibration of non-physical parameters
to control the damage propagation, is generally required.

Y. Shi et al. [109,131] used stress-based criteria and fracture mechanics techniques to
capture composite laminate damage initiation and evolution of damage during an impact event.
The nonlinear shear properties of composites were defined by a shear stress–strain semi-
empirical formula. X-ray radiography was used to validate the proposed numerical model.
Ansari and Chakrabarti [132] conducted a numerical investigation on the penetration and
perforation behaviour of FRP composite laminates under impact loading. The effects of
boundary conditions and thickness-to-span ratio were discussed. Donadon et al.[133,134],
Faggiani et al. [135] and Falzon et al. [136,137] proposed a 3D Computational Damage
Mechanics (CDM) based material damage model to capture the intralaminar degradation of
composite laminates with nonlinear shear behaviour. This model was combined with cohesive
elements to investigate impact damage and captured the permanent indentation caused by low
velocity impact.

Figure 2-17. Out-of-plane displacement magnitude for stiffened composite panel FE model
after impact showing permanent indentation as a result of the damage process (in mm) [135].

Along with the development of the experimental methodology, various numerical


approaches have been employed to model the composites under high-velocity impact events
[138,139]. For example, Nishikawa et al. [138] presented a numerical simulation based on the
Lagrange multiplier method to predict the impact-induced deformation and damage of

34
CHAPTER 2

composites subjected to high-velocity soft-body impacts. In this method, severe contact-


induced deformation caused by a relatively large contact force can be effectively determined
by adjusting the time-scale by using a small time step. Liu et al. [139] used an explicit FE code
to predict damage in a composite leading edge. The Smoothed Particle Hydrodynamic (SPH)
method was used to model the deformation and flow ofthe soft-body gelatine projectile, which
was employed to simulate a bird-strike. In the SPH method, the FE mesh for the impacting
body is replaced by interacting particles, that are able to better represent the deformation and
flow of the projectile over the target. Cantwell et al. [46,48,66,140] have performed a number
of studies on the high-velocity impact responses of CF/Epoxy composites, in which specimens
with various lay-up configurations and thicknesses were evaluated. The energy-absorbing
capability and failure mechanisms of these composites were determined and a model was
developed to predict the degree of perforation of the impacted target specimen. This model
enabled the evaluation of the effects of target thickness, and other factors, on the threshold
impact velocity, and the associated impact energy, for damage initiation and for perforation of
the composite target.

2.7 Summary
This chapter introduced the fundamentals of the composite materials, and gave a review
of relevant past and current research work on composite materials subjected to impact loading.
A brief overview of historical background and classification of composite materials, including
fibre reinformcent, polymer matrix, were given. The fabricaiton processes for thermoplastic
and thermoset composites were introduced, and the effect of fabrication parametres on
mechanical properties of composites were elaborated.

This is followed by a review of the impact performance of composite materials under


low-velocity and high-velocity impact. The impact response and failure mechanism, as well as
the effect of mechanical properties, e.g. strength and stiffness of composite matrix and
reinfrocement, and geometrical factors, e.g. projectile shape and laminate lay-up configuration,
on the overall impact performance of the composite structures, were discuessed. Subsequently,
a number of Non-Destructive Testing (NDT) techniques for damage assessment of composites
were introduced and compared their applications and limits. There has been relatively little
research on comparing the effects of changing the impact velocity and energy levels over a
relatively wider range when using a rigid or soft impactor, such as might arise from debris,

35
CHAPTER 2

birds or tool impacts, to strike a thermoplastic composite and a thermoset composite with the
same fibre architecture.

Finally, this chapter also introduced the fundamental Finite Element (FE) modelling
techniques, including the interlaminar modelling, intralaminar modelling, non-linear shear
modelling and impact behaviour modelling. A number of numerical research work based on
various modelling methods were reviewed, which characterised and predicted the response of
impact and energy-absorption of the composite materials. Furthermore, a number of factors
that affect the impact response of composite structures, including the loading rate, projectile
geometry and material constituents of the composite structures were reviewed, which
contributed to increase the design efficiency and reduce the maintenance costs of composite
structures.

36
Chapter 3. Materials and Experimental
procedures

3.1 Introduction
The carbon fibre reinforced thermoplastic and thermoset composite specimens, used in
this study, were fabricated by employing either a unidirectional-fabric prepreg or a woven-
fabric prepreg. Various properties of these specimens, including fibre type, fibre volume
friction, thickness, toughening agent, dimension, moulding method etc., were introduced in this
chapter. Various types of experiments and technologies, e.g. low-velocity (drop weight) testing
and high-velocity (gas gun) testing were performed to understand the failure mechanisms and
determine the performance of these specimens under impact loading.

3.2 Materials
3.2.1 Unidirectional fibre reinforced composites
The unidirectional AS4 carbon fibre reinforced poly (ether-ether-ketone) (CF/PEEK)
prepregs (APC-2) and the T700 carbon fibre reinforced epoxy (CF/epoxy) prepregs (BA9913)
were employed to fabricate the thermoplastic and thermoset specimens, respectively. The
photographs of these unidirectional CF/PEEK and CF/epoxy specimens are shown in Figure
3-1. Note that specimen categories can be determined visually from external features, as the
pale crystalline polymer by precipitation can be seen on the surface of the CF/PEEK specimens
while the CF/epoxy specimens present good gloss of carbon fibres.

37
CHAPTER 3

Figure 3-1. Photographs of the unidirectional CF/PEEK and CF/epoxy composite


specimens.

In the CF/Epoxy specimens, 20-30 wt% thermoplastic polyether-sulfone toughening


agent was used to toughen the matrix system. Note that CF/PEEK employed AS4 carbon fibres
and the T700 carbon fibres, employed in the thermoplastic and thermoset laminates
respectively, have very similar mechanical properties [141,142], which confirms the facility of
the comparative study. Both the CF/PEEK and CF/epoxy specimens have the same layup of
[903/03]2s and the same overall thickness of 3 mm. The overall density of the CF/PEEK and
CF/Epoxy composite specimens was 1.60 g.cm-3 and 1.62 g.cm-3, respectively. It should be
noted that the objectives here were (a) to achieve virtually the same density for the two types
of composites (i.e. the CF/PEEK and the CF/Epoxy composites) and (b) for the specimens
made from the two types of composites to have very similar values of thickness. The primary
properties of both specimens are summarised in Table 3-1.

38
CHAPTER 3

Table 3-1. Primary properties of the unidirectional CF/PEEK and CF/Epoxy laminates.
CF/PEEK CF/Epoxy
Prepreg APC-2 BA9913
Matrix system Poly(ether-ether ketone) Epoxy (toughened)
Fibre type Unidirectional AS4 Unidirectional T700
Fibre volume fraction 60 % 68 %
Overall density 1.60 g cm-3 1.62 g cm-3
Moulding Press moulding Autoclave consolidation
Lay up [03/903]2s [03/903]2s
Thickness 3 mm 3 mm
Dimensions 150 × 100 mm 150 × 100 mm

3.2.2 Woven-fabric fibre reinforced composites


The woven-fabric CF/PEEK and CF/Epoxy composite specimens were made from the
same T300 carbon fibres but two different matrix systems (Victrex PEEK 150PF and non-
toughened C-M-P epoxy CP009-3, respectively). The photographs of these woven-fabric
CF/PEEK and CF/Epoxy specimens are shown in Figure 3-2. The specimens were machined,
using a waterjet cutter, from large plaques into 140 x 140 mm test specimens.

Figure 3-2. Photographs of the woven-fabric CF/PEEK and CF/Epoxy composite


specimens.

39
CHAPTER 3

Both types of specimen had a nominal thickness of 2 mm. To achieve this thickness,
seven plies of the thermoplastic prepreg, with a nominal 50% volume fraction of fibres, were
used for the CF/PEEK specimens and eight plies of the thermoset prepreg, with a nominal 55%
fibre volume fraction, were used for the CF/Epoxy specimens. The overall density of the
CF/PEEK and CF/Epoxy composite specimens was 1.54 g.cm-3 and 1.57 g.cm-3, respectively.
The primary properties of both specimens are summarised in Table 3-2.

Table 3-2. Primary properties of the woven-fabric CF/PEEK and CF/Epoxy laminates.
CF/PEEK CF/Epoxy
Matrix system Poly(ether-ether ketone) Epoxy (non-toughened)
Fibre type Woven-fabric T300 Woven-fabric T300
Fibre volume fraction 50 % 55 %
Overall density 1.54 g cm-3 1.57 g cm-3
Moulding Press moulding Autoclave consolidation
Thickness 2 mm 2 mm
Dimensions 140 × 140 mm 140 × 140 mm

3.3 Manufacturing processes


3.3.1 Thermoplastic polymer composites
The unidirectional CF/PEEK laminates were fabricated at the University of Nottingham
using a hot press device provided by Mackey Bowley, United Kingdom. Whilst the woven-
fabric CF/PEEK laminates were manufactured by Haufler Ltd., Germany and used same press
moulding process. The hot press device at the University of Nottingham, shown in Figure 3-3
(a), produces a consolidation temperature of up to 410 °C and a contact load of up to 50 tonnes.
This device also employs a water-cooling system which aids cooling the press platens at a
maximum rate of 4±1 °C/minute.

40
CHAPTER 3

(a)

(b)

(c)

Figure 3-3. Photograph of (a) the thermo compressor and the diagram of (b) the consolidation
jig and (c) the cure cycle for the CF/PEEK laminates consolidation.

41
CHAPTER 3

Before the hot press moulding, unidirectional prepregs were cut into square shape by
utility knife and stacked as desired configuration. The individual plies were pre-tacked by using
soldering iron to weld around the periphery of the stack. The tacked prepregs were placed into
the consolidation jig as presented in Figure 3-3 (b). A steel ‘picture frame’ was used to contorl
the thinckness of the consolidated specimens and stop the flow of material. The release agent
and release films were used for easy demoulding. The cure cycle of the CF/PEEK laminates
consolidation is plotted in Figure 3-3 (c). Note that there is no restrictions on the rate of heat
up and cool down.

3.3.2 Thermoset polymer composites


The unidirectional CF/Epoxy laminates were manufactured at the AVIC Composite
Corporation using an autoclave, shown in Figure 3-4(a), provided by Beian, China. Whilst the
woven-fabric CF/Epoxy laminates were manufactured by Haufler Ltd., Germany and followed
the similar autoclave consolidation process. Typical consolidation jig used for this process
which illustrated in Figure 3-4(b). Note that every three layers of prepreg were vacuumed in
sperate bag for 4 hours and overlaid in desired lay-up and kept for over 12 hours before the
consolidation process. The cure cycle of the CF/Epoxy laminates consolidation is plotted in
Figure 3-4 (c).

42
CHAPTER 3

(a)

(b)

(c)
Figure 3-4. Photograph of (a) the autoclave and the diagram of (b) the consolidation jig and
(c) the cure cycle for the CF/Epoxy laminates consolidation.

43
CHAPTER 3

3.4 Overview of experiments


To determine and compare the performance of the thermoplastic and thermoset
laminates under different rates of impact loading, low- and high-velocity impact testing were
performed. 3D Digial Image Correlation (DIC) measurement was employed with gas gun
testing to monitor the out-of-plane displacement of the speicmens during impact. Note that all
experiments were performed in a laboratory equipped with a constant environmental system
holding a temperature of 22 ± 0.5 °C and a humidity of 50 ± 2 %.

3.4.1 Low-velocity (drop weight) testing


The low-velocity impact experiments were performed by using Instron CEAST 9340
drop weight tower, as shown in Figure 3-5, which is designed to deliver up to 405 J of impact
energy. Each composite specimen was placed on a steel picture-frame which had outer
dimensions that matched those of the 150 mm × 100 mm composite specimens and with a 125
mm × 75 mm window cut-out. This assembly was clamped to the base of the drop-weight tower
using four toggle clamps with rubber tips, which prevented slippage of the composite specimen
during the impact test. The load cell, with a data sampling rate of 500 kHz, was located in the
forward section of the impactor.An anti-rebound system as required in the ASTM standard was
employed to catch the impactor in order to prevent it from impacting the specimen a second
time.

This floor standing drop weight tower also equipped with the CEAST DAS 64K data
acquisition system which designed to run the tower and collect experimental data. This data
acquisition system enables the data collection from an instrumented tup insert with built-in
high-frequency piezoelectric sensor. This sensor is particularly ideal to test fairly brittle
materials such as composite materials.

Before testing, a configuration including the overall mass of drop weight, desired
impact energy, testing mode, specimen information and data acuiquisition should be
established in the CeastVIEW software. Based on the known information, this software
calculates the desired velocity upon impact and then moves the impactor to certain height off
the clamped specimen in return for the gravity potential energy required. For testing mode, all
drop weight experiments in this study were carried out in Compression After Impact (CAI)
testing mode which is common for composite materials and a majority of standards. After

44
CHAPTER 3

impact testing, the load/time traces for each test were recorded while the load/displacement
and energy/time traces could be calculated from the data recorded.

Figure 3-5. Set-up for the drop weight (low-velocity) experiments.

3.4.2 High-velocity (gas gun) testing


The air-feed gas gun was employed to preform high-velocity impact experiments. The
facilities for gas gun experiments are shown in Figure 3-6. Helium gas was used to feed a four-
litre pressure vessel which provides air pressure up to 10 bar and was released into to a three-
metre long barrel by a fast-acting pneumatic valve. In order to fire the projectiles or sabots in
different diametres, this barrel can change its inner diametre to 10 mm, 25 mm or 40 mm by
switching the specific size of the inner barrels.

The velocity of the projectile was controlled by changing the pressure of the vessel and
showed high consistency at the same specified pressure. The oscilloscope recorded the time
difference of the projectile passing two pairs of infrared radiation sensors located at the end of
the barrel. Therefore, the velocity of the projectile can be determined from the time difference

45
CHAPTER 3

and the known distance between two pairs of the sensors. Note that the distance from the end
of the barrel to the specimen was 10 cm and any difference in the velocity of the projectile
measured at the exit point from the barrel compared with that upon it impacting the composite
specimen was less than ±1%. A target chamber consists of steel plates and perspex panels to
provide safety gangrantee and the clear view between the specimen fixture and the camera
stand for 3D DIC measurements.

Figure 3-6. Set-up for the gas gun (high-velocity) experiments.

3.4.3 3D Digital Image Correlation (DIC)


A 3D DIC system was employed to monitor the out-of-plane displacements of the
specimens during gas gun impact, which was measured from the deformations of the non-
impacted (i.e. rear) surface of the composite specimen via the random black speckle-pattern.
As the schematic shown in Figure 3-7, two synchronized high-speed cameras were placed
behind the target chamber at a distance of 925 mm from the centre point of the target. The
distance and angle between two cameras were 410 mm and 25°, respectively, which is as
recommended by GOM Ltd., Germany [143]. A pair of 50-mm fixed focal length lenses,
supplied by Nikon, UK, was employed for both cameras. These two cameras were triggered
simultaneously by the signal generated from the infrared sensors and recorded the deformation
of the rear of the composite specimen. Two light-emitting diode lamps were employed to

46
CHAPTER 3

illuminate the rear surface of the CFRP specimens, which were only turned on a few seconds
before each experiment was initiated.

Figure 3-7. Schematic for the high-velocity impact test with 3D DIC system.

Before DIC measuring process, a calibration object should be measured so that the DIC
measuring system is adjusted and its dimensional consistency is ensured. During this process,
the system determines geometrical parametres of each camera, such as postion and orientation,
and calculates the reference point deviation, i.e. intersection error. According to the measuring
volume and distance in the existing set-up, a 120 × 96 mm calibration panel, with 4 mm
diametre reference points, was used as calibration object to be recorded in 13 defined positions,
in order to determine the geometrical parametres of the cameras and the reference point
deviation. A qualified calibration file would be created via the ‘GOM ARAMIS Professional’
software if the orientation between two cameras is within 25 ± 0.3° and the reference point
deviation is under 0.05. This software was employed also for the data processing of the image
series recorded by the high-speed cameras. Note that a new qualified calibration file generated
for each testing day and only be used with the DIC data recorded on same day. For processing
the measured DIC data, the surface components employed a facet size of 19 pixels, with 15
pixels between each displacement point, following the recommendations for the optimum
speckle size [143].

It should be noted that the 3D DIC technique only records surface displacements and
strains but this is useful in recording the overall displacement response of the panel before
damage and can detect surface damage when it occurs.

47
CHAPTER 3

3.4.4 Ultrasonic C-scan


Ultrasonic C-scan technique was employed to examine all CFRP specimens before
experiemtns, in order to avoid that any potential damage occurred during the process of
manufacturing, transportaion and stroage. Further damage assessments, for the specimens after
impact, were conducted. A portable ‘Prisma 16:64 TOFD’ ultrasonic C-scan device with a 5
MHz probe, supplied by Sonatest Ltd., UK, as shown in Figure 3-8, was employed to assess
potential damaged area of the specimens. This phased array probe integrated with a 16 /mm
encoder which is used to provide encoded position of the probe along the scan axis.

Figure 3-8. Portable C-scan system.

To precisely assess potential damage, a configuration is set for each type of specimen
with different matrix systems and thicknesses. In the configuration, A-scan, refers to a one
dimensional amplitude modulation scan, used to determine the output volume of ultrasonic
signal and the gate measuremnts for pulses. A gate measurement with adjustable length is set
to rectify the extent of signal extraction covering the positive waves reflected from the bottom
surface of the specimen or the potential damage. Therefore, a pre-damaged specimen was used
to understand the A-scan profile of a damaged area. During scanning, C-scan maps, presenting
the damge area as a function of the depth through the thickness of the specimen, show the
damage outline supporting the resolution of up to 0.1 mm². Last but not least, water spray used
to apply a layer of water, as contact agent, on the top surface of the specimen to eliminate the
effect of air and ensure the stability of ultrasonic transimission.

48
Chapter 4. Experimental study on
unidirectional thermoplastic and thermoset
laminates under low-velocity and high-velocity
impact loading

4.1 Introduction
The present chapter describes the results from experimental studies on the behaviour of
continuous carbon fibre polymer matrix composites subjected to a relatively low-velocity or
high-velocity impact, using a rigid, metallic impactor. Drop-weight and gas-gun tests are
employed to undertake the low-velocity and high-velocity impact experiments, respectively.
The carbon-fibre composites are based upon a thermoplastic poly(ether-ether ketone) matrix
(termed CF/PEEK) or a thermoset toughened-epoxy matrix (termed CF/Epoxy), which have
the same fibre architecture of a cross-ply [03/903]2s lay-up (detailed in Chapter 3.2.1). The
studies clearly reveal that the CF/PEEK composites exhibit the better impact performance and
that, at the same impact energy of 10.5±0.3 J, the relatively high velocity test of 54.4±1.0 m.s-
1
leads to more damage in the composites than observed from the low-velocity test where the
impactor struck the composite at 2.56 m.s-1.

Note: The work presented in this chapter was previously published in the Journal of Materials Science
as “The behaviour of thermoplastic and thermoset carbon-fibre composites subjected to low-velocity
and high-velocity impacts”.

49
CHAPTER 4

4.2 Test specification and methods


4.2.1 Drop weight testing
The drop-weight impact experiments were performed using an Instron CEAST 9340
drop tower, as detailed in Chapter 3.4.1, with an instrumented stainless-steel impactor having
a hemispherical head with a diameter of 12.7 mm. The drop-weight impact experiments were
conducted at three energy levels, i.e. 4.5 J, 7.5 J and 10.5 J, and three replicate CF/PEEK and
CF/Epoxy specimens were tested at each energy level. The impact energy was varied by
adjusting the height of the impactor. For the impact energies of 4.5, 7.5 and 10.5 J, with the 3.2
kg impactor, then the corresponding impact velocities were 1.68, 2.16 and 2.56 m.s-1,
respectively. No software filtering of the output force-signal was employed and the software
for the instrumented drop-weight test provided the impact load and displacement as a function
of the time-scale of the impact event.

4.2.2 Gas gun testing


The schematic of the set-up for gas-gun experiments is given in Figure 3-7. In the gas-
gun experiments, the specimens were fixed by four toggle clamps on a steel supporting plate
with a 125 × 75 mm opening. This fixture provides the same boundary condition as used in
drop weight testing. Two high-speed cameras (‘Photron AX200’ cameras supplied by Photron,
Japan) recorded the testing process at their maximum rate of 50,000 frames per second at a
resolution of 384 x 256 pixels. For the 3D DIC measuremnts, the target composite specimens
were painted with a matt-white acrylic paint, which was used to avoid excessive light reflection
from the specimens. A random speckle pattern was hand-painted on the white surface using a
matt-black marker to achieve the maximum contrast. The size of the black speckles was
approximately 1.5 mm.

The ∅12.7 mm projectiles, used in the gas-gun experiments, were machined as cylinder
objects with hemispherical noses, which have the same geometry as the ones used in the drop-
weight impact experiments. These projectiles were manufactured using the 7075-T6
aluminium alloy and had a nominal mass of 7.1 ± 0.1 g. The high-density polyethylene (HDPE)
sabot, with an inner diameter of 12.7 mm and an outer diameter of 24.9 mm, were used for
accelerating the projectiles to a desirable velocity. An aluminium alloy stopper, located at the

50
CHAPTER 4

end of the barrel, was employed to separate the projectile from the sabot and secure the sabot
during the gas-gun experiments.

4.2.3 Post-test inspection


In order to assess the damage suffered by the composite specimens, a visual and an
ultrasonic C-scan inspection were undertaken. The latter used the ‘Prisma’ portable C-scan
equipment (Sonatest Ltd, UK) and the instrumented probe had a scanning frequency of 5 MHz.

4.3 Experimental results


4.3.1 Low-velocity drop weight tests
4.3.1.1 Load versus time and load versus displacement results
Figure 4-1 shows the results for the measured load as a function of time for the
CF/PEEK and CF/Epoxy composite specimens impacted at energy levels of 4.5, 7.5 and 10.5
J using the low-velocity drop-weight test. The corresponding impact velocities were 1.68, 2.16
and 2.56 m.s-1, respectively. Clearly very good reproducibility is recorded for the three replicate
composite specimens. Now, several interesting observations may be made from the results
shown in Figure 4-1.

Firstly, considering the CF/PEEK composites, Figure 4-1(a) at an impact energy


(IE)=4.5 J shows the load versus time curves for the three replicate tests and all these curves
reveal that no detectable damage occurred in the CF/PEEK composites, which is confirmed by
the C-scan results as discussed below. The relatively small amplitude, sinusoidal oscillations
on the rising part of these force versus time curves at an IE=4.5 J are indicative of mass-spring
oscillations, as first analysed in detail in [144–146]. However, the load versus time curves
shown in Figure 4-1(a) for a higher impact energy of IE=7.5 J clearly reveal that the three
replicate curves now all exhibit a very major drop in load at 3.4 ± 4% kN, which is indicative
of the initiation of damage in the composite specimens as is confirmed by the C-scan results
below. Likewise, for Fig. 4(a) at an IE=10.5 J, all three replicate curves show a major decrease
in load at 3.1 ± 10% kN, which again is indicative of damage initiation, again as is confirmed
by the C-scan results shown below.

51
CHAPTER 4

(a)

(b)

(c)

IE = 4.5 J IE = 7.5 J IE = 10.5 J


Figure 4-1. Measured load as a function of time curves for the low-velocity drop-weight tests at
impact energies (IE) of 4.5, 7.5 and 10.5 J: (a) CF/PEEK, (b) CF/Epoxy and (c) a comparison of
typical load versus time curves for the CF/PEEK and CF/Epoxy composites.

Secondly, Figure 4-1(b) shows that a similar behaviour exists for the CF/Epoxy
specimens but for this composite damage occurs even at the lowest impact energy of 4.5 J.
Indeed, the loads at which a major decrease occurs are 2.6 ± 3%, 2.6 ± 2% and 2.8 ± 3% kN
for the values of impact energy of 4.5, 7.5 and 10.5 J, respectively. These interpretations of the
load versus time curves are again confirmed by the C-scan results, as discussed below. Also,
as for the CF/PEEK results, at all three impact energy levels there are relatively minor
sinusoidal oscillations associated with mass-spring oscillations prior to a major decrease in the

52
CHAPTER 4

load associated with damage initiation. Thirdly, Figure 4-1(c) compares a typical load versus
time curve for the CF/PEEK composite with one for the CF/Epoxy composite for each impact
energy level and in Table 4-1 values are given for the average loads for damage initiation and
the maximum recorded loads. It is apparent from these results that at the lowest energy level,
i.e. IE=4.5 J, the CF/PEEK composite exhibited a significantly higher maximum load than
recorded for the CF/Epoxy composite. Further, at this impact energy, whilst damage occurred
in the CF/Epoxy composite, no damage was initiated in the CF/PEEK composite specimens
and their stiffness at relatively high loads was therefore not compromised. In contrast, at the
highest impact energy of 10.5 J, the results for the CF/PEEK and CF/Epoxy composites reveal
very similar values of the maximum load for both types of composite, although the load for
damage initiation is somewhat lower for the CF/Epoxy composite. Indeed, the results show
that for the impact energies of both 7.5 J and 10.5 J the load for damage initiation is consistently
somewhat higher for the CF/PEEK than that for the CF/Epoxy composite specimens, but once
damage has been initiated and has propagated the general shape of the load versus time curves
shown in Figure 4-1(c) for the two types of composite are similar.

Table 4-1. Comparison of the measured damage initiation load (i.e. the first major drop in the
measured load versus time curve) and the maximum load obtained from the low-velocity
drop-weight tests. (Note: there was no evidence of a damage initiation load for the CF/PEEK
composite for an impact energy of 4.5 J. The error given is the coefficient of variation from
the replicate experiments).
Type of Damage initiation load (N) Maximum load (N)
composite 4.5 J 7.5 J 10.5 J 4.5 J 7.5 J 10.5 J
CF/PEEK - 3.4±4% 3.1±10% 3.1±7% 3.6±4% 3.9± %
CF/Epoxy 2.6±3% 2.6±2% 2.8±3% 2.8±6% 3.4±2% 3.9 ±2%

Turning to the load as a function of displacement curves shown in Figure 4-2 for the
composite specimens impacted at energy levels of 4.5 J, 7.5 J and 10.5 J, then as may be seen
the maximum out-of-plane displacement increases with increasing impact energy, as would be
expected. No significant difference was found between the values of the maximum
displacements undergone by the CF/PEEK and the CF/Epoxy composite specimens, apart from
possibly at the lowest impact energy of 4.5 J where the CF/Epoxy test specimen underwent a
somewhat higher displacement. This is due to the CF/PEEK composite specimen suffering no
damage at all at this lowest impact energy level, unlike the CF/Epoxy composite. Finally, the

53
CHAPTER 4

intersection point between the load versus displacement curve and the x-axis is the out-of-plane
displacement corresponding to separation between the impactor and the composite specimen,
since for both types of composite the test specimens were never penetrated by the impactor
and a significant rebound of the drop-weight impactor occurred at the end of the impact event.

(a) (b) (c)


Figure 4-2. Measured load as a function of displacement curves for the low-velocity drop-weight
tests for the CF/PEEK and CF/Epoxy specimens at impact energies of (a) 4.5 J, (b) 7.5 J and (c)
10.5 J.

4.3.1.2 Visual and C-scan inspection


For all the composite test specimens, no damage could be detected from a simple visual
inspection of the specimens after the impact test. However, a slight indentation mark was
apparent on the front face of all the test specimens at the impact site.

Figure 4-3 shows the C-scan damage maps obtained from the composite specimens,
impacted at energy levels of 4.5, 7.5 and 10.5 J. At each impact energy, the three replicate
CF/PEEK and three replicate CF/Epoxy composite test specimens were subjected to ultrasonic
C-scanning to identify any interlaminar delamination damage that might have resulted. The
right-hand side scale in Figure 4-3 indicates the location of the delamination damage, as a
function of the depth through the thickness of the specimen, where the dark red colour
represents the front (impacted) face and the dark blue colour represents the rear (non-impacted)
face of the composite specimen. The areal footprint of the damage detected is given in the top
right-hand corner for each specimen and was determined by counting the number of pixels
which had a colour which was not dark blue, as the rear surface simply reflects the ultrasound
and appears as being dark blue in colour.

54
CHAPTER 4

(a)

(b)
Figure 4-3. C-scan damage maps obtained from the (a) CF/PEEK and (b) CF/Epoxy replicate
composite test specimens impacted at a low-velocity (i.e. the drop-weight test) at impact
energies of 4.5 J, 7.5 J and 10.5 J. The corresponding impact velocities were 1.68, 2.16 and
2.56 m.s-1, respectively.

55
CHAPTER 4

In Figure 4-3(a) for the CF/PEEK composites impacted at 4.5 J it is apparent that there
is no delamination damage, as expected from the force versus time curves shown in Figure 4-
1(a). For the CF/PEEK specimens impacted with energies of 7.5 and 10.5 J delaminations are
detected. Such delamination damage would be expected mainly to occur preferentially between
the adjacent 0o and 90o ply directions in the [03/903]2s cross-ply lay-up, since at these interfaces
the values of stiffness, in any one direction, of the adjacent ply layers change dramatically from
one ply to the next. Thus, the interfacial stresses thereby generated tend to initiate
delaminations [147]. Further, the area of the delaminations increases in extent as the impact
energy is increased from 7.5 J to 10.5 J. In contrast, delaminations are observed to occur in the
CF/Epoxy composites at all the impact energies employed and again the damage area steadily
increases in size as the impact energy is increased, see Figure 4-3(b). For both types of
composite, which have the same fibre lay-up of [03/903]2s, the delamination damage that is
furthest from the impact face is a slightly lighter blue in colour than the background and occurs
at a depth of almost 3 mm. This damage is aligned along the longer dimension of the rectangular
specimen, consistent with the direction of the 0o outer-ply layer since, in general, delaminations
propagate in a direction determined by the orientation of the ply beneath the delamination [147].
The values of the average damage areas, i.e. the area of the delamination footprint, measured
from the C-scan maps are summarised in Table 4-2. This comparison clearly shows that the
CF/PEEK composite, when impacted at these relatively low velocities, suffers significantly
less damage than the CF/Epoxy composite when the results are compared at the same energy
level.

Table 4-2. Comparison of the footprint of the damage areas obtained from the low-velocity
drop-weight tests as measured from the C-scans. (The error given is the coefficient of
variation from the replicate experiments.)
Average damage area (mm2)
Type of composite
4.5 J 7.5 J 10.5 J
CF/PEEK 0±0% 274±11% 335±5%
CF/Epoxy 267±6% 395±3% 517±6%

56
CHAPTER 4

4.3.2 High-velocity gas gun impact tests


4.3.2.1 Visual and C-scan inspections
As in the case of the low-velocity drop weight tests discussed above, for all the
composite test specimens no damage could be readily detected from a simple visual inspection
of the specimens after being impacted using the gas gun test. However, a slight indentation
mark was apparent on the front face of all the test specimens at the impact site.

Also, as for the low-velocity drop-weight tests, C-scan damage maps were determined
for all the composite specimens that had been impacted using the high-velocity gas-gun and it
is convenient to first discuss such results. Figure 4-4 shows C-scan damage maps obtained from
the CF/PEEK and CF/Epoxy composite specimens using the gas-gun and impacted at an
average high-velocity of 54.4±1.0 m.s-1 , which gives a corresponding average impact energy
of 10.5±0.3 J. Duplicate impact tests were undertaken for each type of composite, as shown.
The areal footprints of the damage for each test specimen are given in the top-right corner of
the C-scan maps with the respective impact energy given in the top left-hand corner. These
results demonstrate that the CF/PEEK and CF/Epoxy composites exhibit a similar pattern with
values of the average damage area of 1462±37 mm2 and 1898±395 mm2, respectively. Thus,
the CF/Epoxy composite clearly has the higher average damage area but with a significantly
larger degree of scatter.

Figure 4-4. C-scan damage maps obtained from the CF/PEEK and CF/Epoxy duplicate
composite test specimens impacted at a high-velocity (i.e. the gas-gun test). The average
impact velocity and energy were 54.4±1.0 m.s-1 and 10.5±0.3 J, respectively. (The actual
impact energy for each test is shown in the top left-hand corner for each specimen.)

57
CHAPTER 4

4.3.2.2 Out-of-plane displacement results


It is difficult to measure the load versus time relationship in a high-velocity gas-gun
test and such curves can show considerable dynamic effects, as discussed below when the
modelling of such tests is considered. For these reasons, measuring the displacement of the
impacted specimen is typically undertaken. Figures 4-5 and 4-6 show the typical out-of-plane
displacement contours obtained from the DIC measurements with respect to the time-scale of
the impact event for the CF/PEEK and CF/Epoxy test specimens impacted at a velocity of
54.4±1.0 m.s-1 and an impact energy of 10.5±0.3 J. For both composites, it is clear that the
deformation is initially localised around the area of the composite specimen that was struck by
the impactor. However, as the maximum out-of-plane displacement in the central region of the
specimen increases, the deformation then extends elliptically, with the major axis of the ellipse
aligned with the lengthwise dimension of the specimen, which is the direction of the 0o plies.
For an impact energy of 10.5±0.3 J, the maximum out-of-plane displacements of the specimens
are lower in the case of the high-velocity gas-gun test compared with the low-velocity drop-
weight test, as may be seen from comparing Figures 4-2(c), 4-5 and 4-6. Indeed, these results
reveal that the values from the high-velocity test are approximately 3 mm for the CF/PEEK
composite and 3.4 mm for the CF/Epoxy composite, which may be compared with the values
of approximately 5 mm for both types of composite from the low-velocity tests at an impact
energy of 10.5 J. These observations will be discussed later in the context of the modelling
studies in Chapter 5.

58
CHAPTER 4

Figure 4-5. Maps and diagrams of the typical full-field out-of-plane displacements of the rear surface of the composite specimens measured for
both the loading (upper map strips) and unloading (lower map strips) phases for the high-velocity gas-gun experiments at 54.4±1.0 m.s-1 and an
impact energy of 10.5±0.3 J for the CF/PEEK composite test specimens. (For the displacement maps the value of the displacement is given by
the scale bar, with the brightest-red colour representing in the range of about 3.2 to 3.6 mm. For the out-of-plane displacement diagrams the time
interval between the dashed lines is 0.02 ms. The diagrams give the displacement values taken across the middle section of the specimen, as
illustrated schematically in the scaled map to their right.)

59
CHAPTER 4

Figure 4-6. Maps and diagrams of the typical full-field out-of-plane displacements of the rear surface of the composite specimens measured for
both the loading (upper map strips) and unloading (lower map strips) phases for the high-velocity gas-gun experiments at 54.4±1.0 m.s-1 and an
impact energy of 10.5±0.3 J for the CF/Epoxy composite test specimens. (For the displacement maps the value of the displacement is given by
the scale bar, with the brightest-red colour representing in the range of about 3.2 to 3.6 mm. For the out-of-plane displacement diagrams the time
interval between the dashed lines is 0.02 ms. The diagrams give the displacement values taken across the middle section of the specimen, as
illustrated schematically in the scaled map to their right.)

60
CHAPTER 4

4.3.3 Comparison of damage inflicted by the low-velocity and


high-velocity tests
In Table 4-3 results are shown comparing the damage areas, as measured from the C-
scans, for the CF/PEEK and CF/Epoxy composite specimens from both the low-velocity (drop-
weight) and the high-velocity (gas-gun) tests but with a similar impact energy of approximately
10.5 J.

Table 4-3. Comparison of the footprint of the damage areas as measured from the C-scans
obtained from the low-velocity drop-weight and high-velocity gas-gun tests at a similar
impact energy of 10.5±0.3 J. (The error given is the coefficient of variation from the replicate
experiments.)
Damage initiation load (N)
Type of composite
at 2.56 m.s-1 at 54.4±1.0 m.s-1
CF/PEEK 335±5% 1462±3%
CF/Epoxy 517±6% 1898±21%

Firstly, as noted previously, for the low-velocity tests at 2.56 m.s-1 then the CF/Epoxy
composites suffer a significantly greater extent of damage than the CF/PEEK composites.
Secondly, at the higher test velocity of 54.4±1.0 m.s-1 the average damage area for the
CF/Epoxy composites is again greater than for the CF/PEEK composites. Notwithstanding, as
was noted above, the CF/Epoxy composite specimens show a relatively large degree of large
scatter for the values recorded for the damage areas from the C-scan tests, compared with the
far more consistent values for the CF/PEEK composite. Thus, it cannot be definitely stated that
the CF/Epoxy composite exhibited statistically significantly more damage than the CF/PEEK
composite when they were both tested using an impact energy of 10.5±0.3 J, i.e. at the higher
test velocity of 54.4±1.0 m.s-1. Thirdly, both the CF/PEEK and CF/Epoxy composite specimens
clearly exhibit very significantly larger damage areas when subjected to the high-velocity tests
of approximately 54.4±1.0 m.s-1, when compared with the damage areas associated with the
low-velocity tests at 2.56 m.s-1. This observation undoubtedly arises from several causes, since
in the high-velocity tests (a) the CF/PEEK and CF/Epoxy composite specimens have excessive
contact pressure between the projectile, i.e. less time to absorb the impact energy, and therefore
the deformation in the specimen is initially more localised around the impact site with an

61
CHAPTER 4

increased curvature of the specimen local to the impact site, which will induce more extensive
localised delaminations throughout the composite plies, (b) these extensive localised
delaminations will then readily propagate along the various 0o/90o ply interfaces and (c) the
values of the interlaminar and matrix fracture energies will be somewhat lower for both the
CF/PEEK and CF/Epoxy composites due to the relatively high strain-rate effects associated
with the high-velocity test and therefore delaminations and matrix damage will initiate and
evolve more readily in such tests. The overall consequence is that the damage inflicted is more
extensive in the high-velocity tests compared with the low-velocity tests, as indeed is observed
from the greater area of delaminations in the C-scans of the high-velocity tests, see Figures 4-
3 and 4-4, and Table 4-3.

4.4 Summary
The impact behaviour of a thermoplastic CF/PEEK composite and a toughened-
thermoset CF/Epoxy composite, which possessed the same cross-ply lay-up of [03/903]2s, has
been investigated. The relatively low-velocity impact experiments, at impact velocities of 1.68,
2.16 and 2.56 m.s-1 giving impact energies of 4.5, 7.5 and 10.5 J, were undertaken using a drop-
weight test. For the high-velocity experiments, a gas-gun was employed using an impact
velocity of 54.4±1.0 m.s-1 which resulted in an impact energy of 10.5±0.3 J. In all the tests a
rigid, metallic impactor was used with a hemispherical-shaped head. The damage inflicted in
the composites was assessed by both a visual inspection and by ultrasonic C-scanning.

For the low-velocity drop-weight tests, the experimental results demonstrated that, at
the same impact energy, the CF/PEEK composite always possessed a superior impact
performance compared with the CF/Epoxy. This was reflected by the CF/PEEK composite
requiring a higher load to initiate damage and a smaller area of delamination damage being
present in the CF/PEEK composite post-impact.

In the high-velocity tests using the gas-gun, the results revealed that the CF/PEEK and
CF/Epoxy composites exhibited a similar damage pattern with values of the average damage
area of 1462±37 mm2 and 1898±395 mm2, respectively. Thus, the CF/Epoxy composite clearly
exhibits the higher average damage area but with a significantly larger degree of scatter.
Furthermore, these damage areas are significantly greater than those recorded from the low-
velocity drop-weight test, at the same impact energy of 10.5±0.3 J, by a factor of about four
for both types of composite. The out-of-plane displacements of the composite specimens were

62
CHAPTER 4

measured as a function of time during the impact event with the maximum values being
approximately 3 mm for the CF/PEEK composite and 3.4 mm for the CF/Epoxy composite.

Therefore, the present detailed study has revealed the effects of the impact energy and
velocity on the behaviour of a thermoplastic- and a toughened thermoset-matrix CFRP
composite, where both have the same cross-ply fibre-architecture.

63
Chapter 5. Numerical modelling on
unidirectional thermoplastic and thermoset
laminates under low-velocity and high-velocity
impact loading

5.1 Introduction
The present chapter develops a numerical model on the behaviour of continuous
carbon-fibre/polymer matrix composites subjected to a relatively low-velocity or high-velocity
impact, using a rigid, metallic impactor (experiments detailed in Chapter 4). This
computationally-efficient, finite-element model that has been developed is generally successful
in capturing the essential details of the impact test and the impact damage in the composites,
and has been used to predict the loading response of the composites under a high-velocity
impact.

5.2 Composite damage model


5.2.1 Brief overview of the damage model
It is obviously of great interest to attempt to model the experimental result from the
impact tests described in Chapter 4 and, once validated, to use the model to undertake
predictive studies. The impact event for the relatively hard impactor striking the CFRP

Note: The work presented in this chapter was previously published in the Journal of Materials Science
as “The behaviour of thermoplastic and thermoset carbon-fibre composites subjected to low-velocity
and high-velocity impacts”.

64
CHAPTER 5

composite specimen was therefore modelled using the ‘Abaqus/Explicit 2018’ finite-element
(FE) code [148]. An overall flow chart of the basic two-dimensional FE methodology is shown
in Figure 5-1 and the model is shown schematically in Figure 5-2.

Figure 5-1. The implementation of the FE model showing schematically the flow chart, for
one computation time-step, for a single element for modelling the interlaminar and
intralaminar damage

65
CHAPTER 5

Figure 5-2. Schematic of the FE model.

The impactor was modelled as a spherically-shaped, rigid surface, with a reference


lumped mass of 3.2 kg or 7.1 g for the low- and high-velocity tests, respectively. The composite
specimen was defined using continuum shell elements and thus only the in-plane material
properties are required for the numerical modelling. The elements had a size of 1 mm × 1 mm
and this mesh size was found to give mesh-independent results. Cohesive surfaces were defined
at the 0o /90o interfaces to capture the interlaminar damage. A general contact algorithm was
employed to govern the global contact and the friction coefficients for the metal/composite-ply
and composite-ply/composite-ply interfaces were set as 0.2 and 0.25, respectively [149,150].
The computational accuracy was set as ‘double procession’ to reduce the accumulation of error
during the simulation. The FE model for the low-velocity drop-weight test was typically run as
individual time steps, t, of 0.06ms for about 100 steps, which represents the order of the
experimentally measured time of about 6 ms for the complete impact event to take place. Whilst,
for the high-velocity gas-gun test, the model was typically run as individual time steps, t, of
0.005ms for about 100 steps, which represents the order of the experimentally measured time
of about 0.5 ms for the complete impact event to take place. A computational time-step was
performed for every appropriate single element in the FE model. The modelling runs were
stopped when the defined total time for the impact event had expired. Computations were
undertaken using 32 CPUs on a ‘Linux Cluster’ with a run time of 12 to 15 hours. The material
coordinate system for the CFRP laminate was defined with the lengthwise direction of the
specimen being the 11-direction and the widthwise direction being the 22-direction. For the 0o
plies, the fibre direction was aligned with the 11-direction, and for the 90o plies, the fibres were
aligned in the 22-direction.

66
CHAPTER 5

5.2.2 The intralaminar damage model


5.2.2.1 Initiation of intralaminar damage
The model for predicting the initiation of any intralaminar damage was based upon
Hashin’s 2-D theory [151,152] which assumes that the undamaged composite exhibits linear-
elastic behaviour. In Hashin’s damage model, four different types of damage mechanisms,
which arise from tensile fibre failure, compressive fibre failure, tensile matrix failure and
compressive matrix failure, are employed to capture the initiation of intralaminar damage in
the unidirectional-fibre sub-plies. The general forms of the damage criteria in Hashin’s
approach to model the initiation of these four different types of damage are given by:

for tensile fibre failure (𝜎̂11 ≥ 0):

𝜎̂11 2
𝐹𝑓𝑡 = ( 𝑇) (5-1)
𝑋

for compressive fibre failure (𝜎̂11 < 0):

𝜎̂11 2
𝐹𝑓𝑐 = ( 𝐶) (5-2)
𝑋

for tensile matrix failure (𝜎̂22 ≥ 0):

𝜎̂22 2
𝐹𝑚𝑡 = ( 𝑇) (5-3)
𝑌

for compressive matrix failure (𝜎̂22 < 0):

2
𝜎̂22 2 𝑌𝐶 𝜎̂22 𝜏̂12 2
𝐹𝑚𝑐 = ( 𝑇 ) + [( 𝑇 ) − 1] 𝐶 + ( 𝐿 ) (5-4)
2𝑆 2𝑆 𝑌 𝑆

67
CHAPTER 5

In the above equations, the indices on the terms 𝐹𝑓𝑡 , 𝐹𝑓𝑐 , 𝐹𝑚𝑡 and 𝐹𝑚𝑐 represent the four types of
damage of tensile fibre failure, compressive fibre failure, tensile matrix failure and compressive
matrix failure, respectively, and failure is predicted to occur when 𝐹  1. The parameters, 𝑋 𝑇
and 𝑋 𝐶 denote the tensile and compressive strengths in the longitudinal fibre-direction,
respectively. The terms 𝑌 𝑇 and 𝑌 𝐶 are the tensile and compressive strengths in the transverse
direction, respectively; 𝑆 𝐿 and 𝑆 𝑇 = 𝑌 𝐶 /2 denote the shear strengths in the longitudinal and
transverse directions to the fibres, respectively; and the term 𝜎̂11 , 𝜎̂22 and 𝜏̂12 are components
of the effective stress tensor, 𝜎̂, that are used to evaluate the above criteria.

5.2.2.2 The evolution of intralaminar damage


Corresponding to the damage initiation criteria defined in Hashin’s theory, Equations
(5-1) - (5-4), four damage parameters, 𝑑𝑓𝑡 , 𝑑𝑓𝑐 , 𝑑𝑚
𝑡 𝑐
and 𝑑𝑚 , were implemented in the damage
evolution model, which arise from tensile fibre failure, compressive fibre failure, tensile matrix
failure and compressive matrix failure, respectively. These parameters have the value of 0 when
the element is undamaged and 1 when fully damaged. A general form of the damage variable,
d, once a particular damage mechanism initiates, is given by [148]:

𝜀 𝑓 (𝜀 − 𝜀 0 )
𝑑= (5-5)
𝜀(𝜀 𝑓 − 𝜀 0 )

where the strain, 𝜀, is the applied strain and the strain values 𝜀 0 and 𝜀 𝑓 are those corresponding
to the initiation of damage and final failure, respectively. For tensile fibre or compressive fibre
0 𝑓
failure, the terms 𝜀, 𝜀 0 and 𝜀 𝑓 are assigned to be 𝜀 = 𝜀11 , 𝜀 0 = 𝜀11 and 𝜀 𝑓 = 𝜀11, respectively;
with the second superscripts of t or c being used to indicate tensile or compressive stresses,
respectively. Similarly, for tensile or compressive matrix failure, the terms 𝜀, 𝜀 0 and 𝜀 𝑓 are
0 𝑓
assigned to be 𝜀 = 𝜀22 , 𝜀 0 = 𝜀22 and 𝜀 𝑓 = 𝜀22 , respectively; with the second superscripts of t
or c being used to indicate tensile or compressive stresses, respectively.

Now, the applied strain, 𝜀, may be deduced from interrogating the FE output for any
element for a given time step. In the damage evolution model, the values of the initial failure
strains, 𝜀 0 , are equal to the strain values corresponding to damage initiation, which may be

68
CHAPTER 5

directly obtained from the computation via implementing Equations. (5-1) - (5-4). The final
failure strains, 𝜀 𝑓 , are given by [153]:

for tensile fibre failure:

𝑓𝑡
𝜀11 = 2𝐺𝐼𝑐 |𝑓𝑡 /(𝑋 𝑇 𝑙𝑐 ) (5-6)

for compressive fibre failure:

𝑓𝑐
𝜀11 = 2𝐺𝐼𝑐 |𝑓𝑐 /(𝑋 𝐶 𝑙𝑐 ) (5-7)

for tensile matrix failure:

𝑓𝑡
𝜀22 = 2𝐺𝐼𝑐 |𝑚𝑡 /(𝑌 𝑇 𝑙𝑐 ) (5-8)

for compressive matrix failure:

𝑓𝑐
𝜀22 = 2𝐺𝐼𝑐 |𝑚𝑐 /(𝑌 𝐶 𝑙𝑐 ) (5-9)

where the terms 𝐺𝐼𝑐 |𝑓𝑡 and 𝐺𝐼𝑐 |𝑓𝑐 are the tensile and compressive intralaminar ply fracture
energies in the longitudinal fibre-direction, and 𝐺𝐼𝑐 |𝑚𝑡 and 𝐺𝐼𝑐 |𝑚𝑐 are the tensile and
compressive interlaminar ply fracture energies in the transverse to the fibre-direction. The
characteristic length, 𝑙𝑐 is equal to the edge length of the element.

Three damage variables, 𝑑𝑓 , 𝑑𝑚 and 𝑑𝑠 , which reflect fibre damage, matrix damage
and shear damage, respectively, may then be derived from these damage parameters, 𝑑𝑓𝑡 , 𝑑𝑓𝑐 ,
𝑡 𝑐
𝑑𝑚 and 𝑑𝑚 , as follows [148]:

for fibre damage:

69
CHAPTER 5

𝑑𝑓𝑡 , 𝜎̂11 ≥ 0
𝑑𝑓 = { 𝑐 (5-10)
𝑑𝑓 , 𝜎̂11 < 0

for matrix damage:

𝑡
𝑑𝑚 , 𝜎̂22 ≥ 0
𝑑𝑚 = { 𝑐 (5-11)
𝑑𝑚 , 𝜎̂22 < 0

for shear damage:

𝑑𝑠 = 1 − (1 − 𝑑𝑓𝑡 )(1 − 𝑑𝑓𝑐 )(1 − 𝑑𝑚


𝑡 𝑐
)(1 − 𝑑𝑚 ) (5-12)

Now, prior to damage initiation the composite laminate is taken to be linear-elastic,


with the stiffness matrix of a plane-stress orthotropic material. Once the damage initiates and
starts to evolve the response of the material is computed from:

𝜎 = 𝐶𝑑 𝜀 (5-13)

where 𝐶𝑑 is the damaged elasticity matrix, which has the form [148]:

(1 − 𝑑𝑓 )𝐸11 (1 − 𝑑𝑓 )(1 − 𝑑𝑚 )𝜈21 𝐸11 0


1
𝐶𝑑 = [(1 − 𝑑𝑓 )(1 − 𝑑𝑚 )𝜈12 𝐸22 (1 − 𝑑𝑚 )𝐸22 0 ] (5-14)
𝐷
0 0 (1 − 𝑑𝑠 )𝐺12 𝐷

where 𝐷 = 1 − (1 − 𝑑𝑓 )(1 − 𝑑𝑚 )𝜈12 𝜈21 , and 𝐸𝑖𝑖 (𝑖 = 1,2) is the elastic modulus in the
longitudinal or transverse directions, G12 is the shear modulus and 𝜈𝑖𝑗 (𝑖, 𝑗 = 1,2 𝑖 ≠ 𝑗) are the
Poisson's ratios. Now, the values of the damage variables, 𝑑𝑓 , 𝑑𝑚 and 𝑑𝑠 , reflect the current
state of fibre damage, matrix damage and shear damage, respectively, and may be calculated
70
CHAPTER 5

from Equations (5-10) to (5-12). Thus, the degraded stresses acting in any element for any time
step for an applied strain,  , can now be computed from Equations (5-13) and (5-14). These
degraded stresses and strains may then be updated, as being the ‘new model state’, for a given
element in the next time step of the run of the FE model, see Figure 5-1. For the simulations of
the extent of intralaminar damage as a function of the time-scale of the impact event that are
deduced from the model then, following earlier work [135], a value for the damage parameter
of equal to, or greater, than 0.9 is used to define the relatively intense intralaminar damage, i.e.
to calculate the areas indicated by a red colour in the figure shown later. This value is
determined based on a number of attempts from previous work to capture sufficient
intralaminar damage without interfusing excessive element distortion.

From the above discussions, it is evident that if any of the four intralaminar damage
mechanisms have been activated as determined from the Hashin criteria, see Equations (5-1) -
(5-4), then the elastic properties in those elements in the FE model will start to degrade, as
defined above in Equation (5-14), which is based upon the damage parameters calculated in
Equations (5-6) to (5-12). Finally, when any of these three damage parameters meet the
condition, as stated in Equation (15) below, then these damaged elements are deleted from the
model [148]:

𝑑𝑓 > 0.99
𝐷𝑒𝑙𝑒𝑡𝑒 𝑒𝑙𝑒𝑚𝑒𝑛𝑡 𝑖𝑓: {𝑑𝑚 > 0.99 (5-15)
𝑑𝑠 > 0.99

5.2.3 The interlaminar damage model


Interlaminar damage typically involves the initiation and growth of delamination
between the plies that make up the composite laminate and this was captured using the Abaqus
built-in surface-based cohesive (i.e. interface) element using an energy-release approach [148].
The interface element was described via a cohesive (i.e. damage) surface law [118,131,154–
156] where the traction, , is a function of the displacement, , as shown schematically in
Figure 5-3, and is in the form of a bilinear cohesive law for a linear-softening material model.
This damage law is divided into two steps. Before the initiation of any delamination, the
relationship possesses linear-elastic behaviour. Once the damage criterion is satisfied, at a
value of the displacement of 𝛿 𝑜 , the cohesive stiffness degrades linearly until separation of the

71
CHAPTER 5

interface, i.e. the delamination now propagates when the maximum failure displacement of 𝛿 𝑓
is attained. The energy under the bilinear cohesive law is equivalent to the interlaminar fracture
energy, 𝐺𝑐 , and the evaluation of this term, and of the initial stiffness, k, of the interface element,
are discussed below. Finally, the initiation and growth of any intralaminar damage, as discussed
above, significantly influences the extent of interlaminar damage and hence these two damage
modes were modelled to be interactive in the Abaqus simulation, see Figure 5-1.

Figure 5-3. Schematic of the bilinear cohesive surface law.

5.2.3.1 The initiation of interlaminar damage


To determine when interlaminar damage initiates, a quadratic-stress criterion was again
implemented within the Abaqus FE code, and is given by [148,157]:

2 2 2
〈𝑛 〉 𝑠 𝑡
( 0 ) + ( 0) + ( 0)  1 (5-16)
𝑛 𝑠 𝑡

where 𝑛 represents the current stress that is acting normal to the ply, and 𝑠 and 𝑡 represent
the current shear stresses that are acting on the ply. The values of the cohesive strengths, 0𝑛 ,
0𝑠 and 0𝑡 denote the interface shear and tensile strengths, respectively. Employing the
quadratic-stress criterion, the value of the damage initiation displacement, 𝛿 𝑜 , see Figure 5-3,
may be calculated.

72
CHAPTER 5

5.2.3.2 The evolution of interlaminar damage


The embedded cohesive surface law, see Figure 5-3, requires a value of the interlaminar
fracture energy, 𝐺𝑐 , and this represents the area under the bilinear law. The energy-based
Benzeggagh-Kenane (B-K) [84,148,158] mixed-mode propagation criterion was used to derive
a value 𝐺𝑐 for the growth of the delamination between the composite plies, as given by:

𝜼𝑩𝑲
𝐺𝐼𝐼
𝐺𝑐 = 𝐺𝐼𝑐 + (𝐺𝐼𝐼𝑐 − 𝐺𝐼𝑐 )  (5-17)
𝐺𝐼 + 𝐺𝐼𝐼

where 𝐺𝐼𝑐 is the Mode I (opening tensile) interlaminar facture energy, 𝐺𝐼𝐼𝑐 is the Mode II (in-
plane shear) interlaminar facture energy and BK is the B-K Mixed-mode interaction exponent.
The values of all these terms may be experimentally measured [83,84,158,159]. The parameters
𝐺𝐼 and 𝐺𝐼𝐼 are the current Mode I and Mode II energy-release rates, respectively, as calculated
from the FE code by multiplying the relevant local stress by its conjugate displacement. The
stiffness, k, of the cohesive law is given by [155]:

𝛼𝐸22
𝑘= (5-18)
𝑡𝑒

where  is a constant much larger than unity, i.e. 𝛼 ≫ 1, and 𝑡𝑒 is the thickness of an adjacent
ply. From the known values of k, 𝛿 𝑜 and 𝐺𝑐 the cohesive surface law, as shown in Figure 5-3,
may now be completely and quantitatively described. Finally, complete fracture of the interface
element is assumed to occur, and delamination results, when the cohesive traction vanishes at
the end of the degradation step. That is when the displacement, 𝛿, of the interface element, as
determined in the FE code, attains the criterion:

𝛿 ≥ 𝛿𝑓 (5-19)

where 𝛿 𝑓 is the displacement of the element at failure, see Figure 5-3. For the simulations of
the location and extent of interlaminar damage as a function of the impact velocity and energy,

73
CHAPTER 5

and the time-scale of the impact event, that are deduced from the model then, following earlier
work [135], a value corresponding to the displacement ratio of 𝛿/𝛿 𝑓 of equal to, or greater, than
0.9 is used, i.e. to calculate the areas shown by a red colour in the figures shown later.

5.3 Modelling results


5.3.1 Material properties
The material properties of the CF/PEEK and CF/epoxy composites required for the FE
modelling studies are given in Table 5-1. These properties were typically measured at strain-
rates in the range of 10-3 s-1 to 100 s-1.

5.3.2 Simulations of the low-velocity drop weight impact test


results
As for the experimental tests, simulations were undertaken at three different energy
levels of 4.5, 7.5 and 10.5 J, which correspond to the impact velocities of 1.68, 2.16 and 2.56
m.s-1, respectively. The strain-rate for the composite test specimens for the low-velocity drop-
weight tests is about 3 s-1 for an impact velocity of 2.56 m.s-1. It was considered that this still
relatively low value of strain-rate would not lead to significantly different values of the material
properties to those shown in Table 5-1 and these properties were therefore used in the FE
modelling studies for the low-velocity impact tests.

74
CHAPTER 5

Table 5-1. Input properties for the FE modelling studies of the CF/PEEK and CF/Epoxy composites [133,141,147,153,160–164].
Property Unidirectional CF/PEEK sub-ply Unidirectional CF/epoxy sub-ply
Moduli (GPa) 𝐸11 = 127; 𝐸22 = 10.3; 𝐺12 = 5.7 𝐸11 = 130; 𝐸22 = 7.7; 𝐺12 = 4.68
Poisson’s ratio 𝜈12 = 0.3 𝜈12 = 0.3

𝑋 𝑇 = 2070; 𝑌 𝑇 = 85 𝑋 𝑇 = 1950; 𝑌 𝑇 = 75
Strength values (MPa) 𝑋 𝐶 = 1360; 𝑌 𝐶 = 276 𝑋 𝐶 = 1015; 𝑌 𝐶 = 220
𝑆 𝐿 = 186 𝑆12 = 150

𝐺𝐼𝑐 |𝑓𝑡 = 218; 𝐺𝐼𝑐 |𝑓𝑐 = 104 𝐺𝐼𝑐 |𝑓𝑡 = 133; 𝐺𝐼𝑐 |𝑓𝑐 = 40.0
Ply fracture energies (kJ.m-2)
𝐺𝐼𝑐 |𝑚𝑡 = 1.7; 𝐺𝐼𝑐 |𝑚𝑐 = 2.0 𝐺𝐼𝑐 |𝑚𝑡 = 0.5; 𝐺𝐼𝑐 |𝑚𝑐 = 1.6
Interlaminar fracture energies (kJ.m-2) 𝐺𝐼𝑐 = 1.7 ; 𝐺𝐼𝐼𝑐 = 2.0 𝐺𝐼𝑐 = 0.5 ; 𝐺𝐼𝐼𝑐 = 1.6
Benzeggagh-Kenane coefficient mode-mix
𝜂𝐵𝐾 = 1.09 𝜂𝐵𝐾 = 1.45
exponent
Cohesive strength (MPa) 0𝑛 = 43.0; 0𝑠 = 0𝑡 = 50.0 0𝑛 = 43.0; 0𝑠 = 0𝑡 = 50.0
Cohesive law stiffness (MPa.mm-1) 𝑘 = 6.4 × 105 𝑘 = 6.4 × 105

75
CHAPTER 5

The predicted damage maps from the modelling studies for the three different impact
energies are compared with the corresponding experimental results in Figures 5-4 and 5-5 for
the CF/PEEK and CF/Epoxy composites, respectively. (The experimentally measured maps
which are shown represent the mean value of the damage area that was measured from the
replicate tests, see Figure 4-3 in Chapter 4.)

Figure 5-4. Experimental and simulated (red-coloured) footprints of the damage area for
the CF/PEEK composite specimens for impact energies of 4.5, 7.5 and 10.5 J for the low-
velocity drop-weight impact test. (See Figure 4-3 in Chapter 4 for the explanation of the
experimental C-scan depth scale, etc.)

Figure 5-5. Experimental and simulated (red-coloured) footprints of the damage area for
the CF/Epoxy composite specimens for impact energies of 4.5, 7.5 and 10.5 J for the low-
velocity drop-weight impact test. (See Figure 4-3 in Chapter 4 for the explanation of the
experimental C-scan depth scale, etc.)

76
CHAPTER 5

As may be seen, the increase in the extent of damage as the impact energy is increased
for a given composite is in good agreement with the damage maps that were experimentally
measured using the C-scan test. The general shapes and the locations of the damage areas from
the modelling studies are also in good agreement with the experimental results. The one
exception to the good agreement between the modelling and experimental results is the
prediction of damage occurring in the CF/PEEK composite specimen at an impact energy of
4.5 J, when none was experimentally observed, see Figure 5-4. This was most likely due to the
predicted stress required to initiate such damage, as deduced from Equation (5-16), at the
lowest impact energy used for the CF/PEEK composite being somewhat lower than that
actually required. Hence, interlaminar damage is predicted when none is actually observed
experimentally at the lowest values of impact velocity and energy used in the present study.
Further, as may be seen, the FE model lacks the fidelity to reproduce the exact shape of the
experimentally measured delaminations. Nevertheless, the quantitative predictions of the
damage areas obtained from the FE simulation, demonstrate that the proposed model has the
capability to predict reasonably well the overall area of damage due to a low-velocity impact.

The modelling results for the typical evolution of interlaminar (i.e. delamination)
damage in the CF/PEEK composite specimens subjected to the drop-weight impact test is
presented in the time-scale series of images in Figure 5-6(a) as a series of cross-section images
obtained from the simulations at an impact velocity and energy of 2.56 m.s-1 and 10.5 J,
respectively. The damage evolution of intralaminar (i.e. matrix) damage in the CF/PEEK
composite specimen for the same test conditions is shown in Figure 5-6(b). The intralaminar
matrix damage, such as matrix cracking, arises from transverse stresses acting perpendicular
to the fibres in the plies and there were no predictions of fibre-failure from the modelling
studies during either the low-velocity, or the high-velocity, impact tests. Therefore, it appears
that the stresses induced in the plies are insufficiently high to cause fibre-failure, which is in
agreement with the experimental observations of the tested composite specimens. Finally, an
interesting point to note is that the intralaminar matrix damage under the impactor is much
more localised than the corresponding delamination damage.

77
CHAPTER 5

(a)

(b)
Figure 5-6. Simulations of the (a) interlaminar delamination (red-coloured) damage and (b) intralaminar matrix (red-coloured) damage as a
function of time-scale for the CF/PEEK composites for low-velocity drop-weight tests at an impact energy and velocity of 10.5 J and 2.56
m.s-1, respectively.

78
CHAPTER 5

The load versus time curves obtained from the modelling studies and the experimental
tests for the CF/PEEK and CF/Epoxy composites are compared in Figures 5-7 and 5-8,
respectively. A very good agreement is obtained between the experimental and modelling
results for all three energy levels for both types of composite. Indeed, the modelling results
even capture the small amplitude, sinusoidal oscillations on the rising part of load versus time
curves which are indicative of mass-spring oscillations [144–146], as discussed earlier.

(a) (b) (c)


Figure 5-7. Experimental and simulated load versus time curves for the low-velocity drop-weight
tests for the CF/PEEK composite specimens for: (a) 4.5, (b) 7.5 and (c) 10.5 J impact energy levels.

(a) (b) (c)


Figure 5-8. Experimental and simulated load versus time curves for the low-velocity drop-weight
tests for the CF/Epoxy composite specimens for: (a) 4.5, (b) 7.5 and (c) 10.5 J impact energy levels.

The key experimental and numerical modelling results are compared in Table 5-2 and
these results highlight the good agreement for both types of composite. Indeed, the modelling

79
CHAPTER 5

studies clearly agree with the experimental observations that the CF/PEEK composite, when
impacted at these relatively low velocities, suffers significantly less damage than the CF/Epoxy
composite when the results are compared at the same energy level.

Table 5-2. Comparison of the experimental and numerically-simulated results obtained from
the low-velocity drop-weight tests at an impact energy and velocity of 10.5 J and 2.56 m.s-1,
respectively. (The error given is the coefficient of variation from the replicate experiments.)
CF/PEEK CF/Epoxy
Items
Experiment Simulation Experiment Simulation

Maximum load (kN) 3.9±3% 4.3 3.9±2% 4.2

Average damage area (mm2) 335±5% 380 517±6% 580

5.3.3 Simulations of the high-velocity gas gun impact test results


Many researchers, for example [165–167], have shown that there can be a reduction in
the values of the interlaminar fracture energies for composite materials of up to about 20%
when the strain rate is increased from quasi-static test conditions (i.e. about 10-3 to 100 s-1) to
high-velocity impact velocities of about 55 m.s-1, which in the present tests correspond to a
strain-rate of about 300 s-1. (It was possible to extract from the DIC experimental results the
maximum strain experienced by the composite specimens and, from knowing the time-scale of
the impact event, this information gives a strain rate of about 250 s-1 for the CF/PEEK and
about 350 s-1 for the CF/Epoxy composite test specimens when struck by a rigid impactor at a
velocity of about 55 m.s-1.) As no fibre breakage was observed, or predicted, in the impacted
composite specimens, only the interlaminar and matrix-ply fracture energies, for both Mode I
and Mode II failure, were decreased in the present models of the high-velocity impact tests.
Now, from the literature [165–167] on the effects of strain-rate on the properties of composite
laminates, it is considered that for these high-velocity tests a reduction in the values of the
interlaminar and matrix fracture energies of about 20% should be made to account for this
strain-rate effect. Thus, two different sets of values for the values of the interlaminar and
matrix-ply fracture energies were considered for use in the simulations, i.e. 100% and 80% of
the quasi-static values of 𝐺𝐼𝑐 |𝑚𝑡 , 𝐺𝐼𝑐 |𝑚𝑐 , 𝐺𝐼𝑐 and 𝐺𝐼𝐼𝑐 given in Table 5-1.

80
CHAPTER 5

In the simulations of the gas-gun impact tests an impact velocity of 54.4 m.s-1 was
employed which gives an impact energy of 10.5 J. The simulations of the gas-gun tests were
performed for both the CF/PEEK and the CF/Epoxy composite specimens. The simulation
results were obtained from the numerical FE model with the two different levels of fracture
energy values, as noted above. The predicted damage footprints are shown in Figure 5-9,
together with the experimentally measured C-scan damage maps. The modelling results
accurately predict that (a) the damage areas in the composites are always significantly higher
from the high-velocity gas-gun tests than from the low-velocity drop-weight tests, at the same
impact energy of 10.5 J, see Figures 5-5 and 5-9, and (b) the CF/Epoxy composite undergoes
somewhat greater damage compared to the CF/PEEK composite. However, the results shown
in Figure 5-9 reveal that the areas of the damage footprints predicted by the FE model using
the quasi-static fracture energies are much smaller than those obtained from the experiments.
Obviously, as expected, the models using the reduced values of 80% of the fracture energies
give larger footprints of the damage area and the simulation results are now in somewhat better
agreement with the experimental results. However, the simulations shown in Figure 5-9 reveal
that, even when using 80% of the quasi-static fracture energies, the footprints of the predicted
damage areas are always lower than the values measured experimentally from the C-scans tests.
This is likely to arise from several factors. Firstly, in the high-velocity model simulations, only
the interlaminar and matrix-ply fracture energies were reduced to 80% of their quasi-static
values and the other material properties required for the model were kept at their quasi-static
values. This was done since the fracture energies have been shown to decrease at relatively
high strain-rates [165–167] and the fibre-dominated properties would certainly be expected to
show relatively little change as the strain rate is increased. Secondly, without a detailed
knowledge of the strain-rate effects for the other material properties, it was considered best to
keep all the other material properties at their quasi-static values for consistency. Thirdly, in the
model, the presence of intralaminar-matrix fracture damage is considered to lead to a reduction
in stiffness of the elasticity matrix, i.e. the term 𝐶𝑑 as defined by implementing Equations (5-
10) to (5-14), which represents damage of the ply elements. Thus, intralaminar matrix cracks,
as such, are not physically present in this ‘smeared-crack’ model but only their effects on the
ply element stiffnesses. However, in the experiments, intralaminar matrix cracks could be
diverted at the ply interfaces to then generate more interlaminar delaminations and this may
aspect also account for the extra extent of delamination that was experimentally observed.

81
CHAPTER 5

Figure 5-9. Experimental and simulated (red-coloured) footprints of the damage areas for
the CF/PEEK and CF/Epoxy composite specimens from the high-velocity gas-gun test
using an impact energy and velocity of 10.5±0.3 J and 54.4±1.0 m.s-1, respectively. (See
Figure 4-3 in Chapter 4 for the explanation of the experimental C-scan depth scale, etc.
Results are shown for both 100% and 80% of the quasi-static interlaminar and matrix
fracture energies.)

The FE model, using the 80% of the values of the quasi-static fracture energies, was
also employed to predict the values of the out-of-plane displacement and the results are
compared to the experimental values from the DIC tests, as shown in Figure 5-10. As may be
seen, good agreement is obtained between the experimental and simulation results.
Furthermore, the predicted maximum out-of-plane displacements of 3 mm for the CF/PEEK
and 3.5 mm for the CF/Epoxy composite are in very good agreement with the maximum values

82
CHAPTER 5

of 3 mm and 3.4 mm, respectively, that were experimentally measured, see Figures 4-5 and 4-
6 in Chpater 4. It should be noted that the deflection of the specimen by the impactor is a
structural response of the specimen to the loading by the impactor and, as such, it would not be
expected to be strongly affected by the reduction of the quasi-static fracture energies. To
confirm this, the out-of-plane displacements were also determined at 100% of the quasi-static
fracture energies and it was indeed found that the displacements of the composite specimens
are virtually identical to those predicted when using when using only 80% of the quasi-static
fracture energies in the model. Thus, although the composite specimens are damaged by the
delaminations that occurred, their overall structural responses are not dramatically modified,
as is indeed confirmed by these modelling results.

83
CHAPTER 5

(a)

(b)
Figure 5-10. The experimental and simulated out-of-plane displacement maps obtained from (a) the CF/PEEK and (b) the CF/Epoxy
composite specimens when subjected to the high-velocity gas-gun test at an impact energy and velocity of 10.5±0.3 J and 54.4±1.0 m.s-1,
respectively.

84
CHAPTER 5

For these high-velocity impact tests, the load versus time curves of the CF/PEEK and
CF/Epoxy composites could not be experimentally measured, as discussed earlier. However,
they may now be predicted using the FE model, as shown in Figure 5-11. Interestingly, the
load versus time curves are not very sensitive to the effect of the percentage values of the
fracture energies that are used, since the load versus time response is mostly a dynamic effect.
The presence of the two major peaks in Figures 5-11(a) and (b) is due to the initial rapid
acceleration of the specimen upon being struck by the impactor [168,169]. This rapid loading
leads to a temporary loss of contact between the impactor and the composite specimen which
is then followed by contact being re-established when there is a further phase of loading. The
load versus time impulse (i.e. the area under load versus time curve) is about 0.6 N.s which
will be responsible for the total change in momentum of the incoming impactor. The total
change in momentum would therefore be expected to be 0.6 kg.m.s-1 during the impact event.
The incoming momentum of the impactor is 0.38 kg.m.s-1 and so it would be expected that the
impactor would rebound with momentum of 0.22 kg.m.s-1. This implies that the rebound
velocity of the impactor at the end of the impact event would be about 60% of the impact
velocity, which was not measured but this is consistent with the significant rebound that was
observed. Thus, the gas-gun, and also the drop-weight, impact events were reasonably elastic
in nature with a high proportion of kinetic energy being returned to the impactor.

(a) (b)
Figure 5-11. Load versus time curves as predicted from the FE model for the CF/PEEK and
CF/Epoxy composite specimens subjected to a high-velocity gas-gun test at an impact energy
and velocity 10.5 J and 54.4 m.s-1, respectively. (Results are shown for using 100% and 80%
of the quasi-static interlaminar and matrix fracture energies (FrEn)).
85
CHAPTER 5

5.4 Summary
In the present study, an elastic, two-dimensional finite-element (FE) model has been
developed to simulate the experimental results obtained from both the low-velocity and high-
velocity impact tests in Chapter 4. The model, which is relatively computationally efficient,
has been shown to simulate (a) the loading responses of the composites by the impact event
and (b) the interlaminar and intralaminar damage induced.

For the low-velocity drop-weight tests, the experimental results demonstrated that, at
the same impact energy, the CF/PEEK composite always possessed a superior impact
performance compared with the CF/Epoxy. The results from the FE model for the load versus
time curves from all the tests were in very good agreement with the experimental measurements
and even captured the small amplitude, sinusoidal oscillations on the rising part of the load
versus time curves which are indicative of mass-spring oscillations. Considering the FE
modelling of the high-velocity tests, the simulations have accurately predicted (a) the shape
and values of the out-of-plane displacements as a function of the time-scale of the impact event,
(b) that the damage areas in the composites were always significantly higher from the high-
velocity gas-gun tests than from the low-velocity drop-weight tests, at a comparable impact
energy of 10.5 J, and (c) that the CF/Epoxy composite suffered somewhat greater damage
compared to the CF/PEEK composite. However, the actual footprints of the damage areas
predicted from the modelling studies for the high-velocity tests were always somewhat lower
than the values measured experimentally from the C-scans tests, and reasons for this have been
proposed. For these high-velocity impact tests, the load versus time curves of the CF/PEEK
and CF/Epoxy composites could not be experimentally measured, nevertheless they were
predicted using the FE model. The presence of two major peaks in the load versus time curves
has been identified and ascribed to the initial rapid acceleration of the specimen upon being
struck by the impactor. This rapid loading leads to a temporary loss of contact between the
impactor and the composite specimen which is then followed by contact being re-established
when there is a further phase of loading, before the impactor finally rebounds at the end of the
impact event at about 60% of its impact velocity. Therefore, the numerical FE model has been
developed which is relatively computationally efficient. The results from this model has been
quantitatively validated against the impact response of the composites and captures the
essential aspects of their impact behaviour and can be used (a) to simulate aspects of the high-
velocity tests that cannot be directly measured and (b) to optimise the impact performance of
such materials in industrial applications.

86
Chapter 6. Experimental study on woven-fabric
thermoplastic and thermoset laminates under
high-velocity soft and hard impact loading

6.1 Introduction
The present chapter investigates the impact performance of woven-fabric carbon fibre
composites based upon both thermoplastic- and thermoset-matrix polymers under high-
velocity impact loading by conducting gas gun experiments at impact velocities of up to 100
m.s-1. The carbon fibre reinforced polymers (CFRPs) are impacted using soft (i.e. gelatine) and
hard (i.e. aluminium-alloy) projectiles to simulate either a soft bird-strike or a hard foreign-
body impact (e.g. runway debris), respectively, on composites employed in civil aircraft. The
out-of-plane displacement of the impacted composite specimen is obtained by means of a three-
dimensional Digital Image Correlation (DIC) system for the soft projectile impact on the
composites and the extent of damage is assessed both visually and by using portable C-scan
equipment. The perforation resistance and energy absorbing capability of the composites are
also studied by performing high-velocity impact experiments using the hard-projectile and the
resulting extent and type of damage are identified.

Note: The work presented in this chapter was previously published in the Applied Composite Materials
as “The impact performance of woven-fabric thermoplastic and thermoset composites subjected to high-
velocity soft- and hard-impact loading”.

87
CHAPTER 6

6.2 Test specification and methods


6.2.1 Test specimens and projectiles
The woven-fabric CF/PEEK and CF/Epoxy specimens employed in this study were
detailed in Chapter 3.2.2. The 3D DIC system was employed in the soft-projectile impact
experiments to record the deformation of the CF/PEEK and CF/Epoxy specimens. Therefore,
the non-impacted surface (i.e. the rear surface of the sample) of the CFRP specimens were
prepared according to the instructions of the manufacturer of the DIC system [143] which is
detailed in Chapter 3.4.3. Prior to undertaking the experiments, all the specimens were kept
under a temperature of 22±0.5 °C and a humidity of 50±2 % for 24 hours.

(a)

(b)

(c)

Figure 6-1. Projectiles and sabot for soft- and hard-impact tests: (a) flat-fronted soft
gelatine projectile and (b) HDPE sabot and (c) hemi-spherical head of the hard aluminium-
alloy projectile. (Dimensions are in mm).

88
CHAPTER 6

In order to simulate the impact of a bird-strike and runway debris on the composites,
gelatine projectiles (to represent a relatively soft projectile, e.g. a bird-strike) and aluminium-
alloy projectiles (to represent a hard projectile, e.g. runway debris) were used in the gas-gun
experiments, as shown in Figure 6-1. New projectiles were employed for each impact test.

Now gelatine has been identified as a good substitute material to simulate birds to
overcome the drawbacks of using real birds for the impact tests, as discussed above. The
gelatine projectile generates a similar level of initial shock pressure as that experienced in a
typical bird-strike, which is important in determining the extent and type of local damage that
results in the impacted composite [170]. The gelatine projectiles employed were based on the
bird model developed by Wilbeck and Rand [170].

A well-defined process, which is relatively simple and controllable, has been developed
for preparing the soft-gelatine projectiles to a uniform standard. The ingredients used to prepare
these projectiles were gelatine powder and distilled water. The gelatine powder was supplied
by Honeywell Specialty, Germany. The detailed preparation procedure of the gelatine
projectiles is introduced in Table 6-1. These 3 wt.% gelatine projectiles had a nominal mass of
20 g and possessed a Shore hardness A of 0 to 5 and this enabled them to withstand the inertial
loads during being launched from the gas gun. It is important to note that the shape of the
gelatine projectile from being fired to impacting the target remained essentially unchanged, i.e.
no stretching or compressing of the projectile occurred. To achieve this a sabot, 40 mm in
diameter, was used to provide the projectile acceleration that was required and to reduce the
friction between the soft-projectile and the barrel of the gas-gun.

89
CHAPTER 6

Table 6-1. Detailed preparation procedure of the gelatine projectiles.


Steps Operations
Raise the temperature of the distilled water to 80°C and maintain this
I temperature using a water bath with a thermocouple to monitor the temperature
of the distilled water.
Mix the gelatine powder with distilled water at a specific mass ratio of 3:100,
II and then stir the mixture, using a magnetic stirrer at a stirring rate of 5
revolutions per second, until the gelatine powder is completely dissolved.
Transfer the solution to a beaker and let the solution cool down to room
III
temperature.
Transfer the solution at room temperature to a poly(tetrafluoroethylene) (PTFE)
IV cylindrical mould, which has a paraffin-oil coated onto its surface to prevent
leakage and facilitate the subsequent removal of the solid gelatine projectile.
Seal the mould with cling film placed over the top of the mould to prevent
V
dehydration.
Place the sealed mould into an environmental chamber where the temperature
VI
is kept between 4 to 6°C for at least 12 hours.
Carefully push the solid gelatine cylinder out from the mould and use a
VII digitally-controlled disc saw to cut the gelatine cylinders to the required length
for the gelatine projectiles.

The hard aluminium-alloy projectiles were machined from 7075-T6 aluminium-alloy.


These gave a hard-impact event, for example as might be experienced by impact from runway
debris. The aluminium-alloy projectiles had a nominal mass of 28 g and a nominal diameter of
24.9 mm, which is slightly smaller than the diameter of 25 mm of the barrel of the gas gun.
This was necessary in order for the aluminium-alloy projectiles to achieve a reproducible
maximum velocity.

6.2.2 Gas gun testing


Two different set-ups of the gas gun were used to assess the dynamic deformation and
ballistic response of the composites dependent upon whether the high-velocity impact was via
the relatively soft- or the hard-projectile. Both set-ups fixed the sepcimens between two steel

90
CHAPTER 6

supporting plates, both with a 70 × 70 mm opening, and twelve M8 bolts were used to fasten
the steel plates together. A schematic drawing of the steel plates and the specimen is shown in
Figure 6-2.

Figure 6-2. Schematic of the support for the composite test specimen.
(Dimensions are in mm.)

For ‘Set-up I’ as mentioned in the section 3.4.3, for the impact by the soft-gelatine
projectile, it employed a pair of high-speed cameras, i.e. ‘Phantom Miro M/R/LC310’ cameras
supplied by Vision Research Phantom, USA, to track the deformation history of the rear surface
of the target specimen and record at their maximum rate of 39,000 frames per second. In order
to monitor the full-field deformation of the specimens, the image size was selected as 256 x
256 pixels.

For ‘Set-up II’, the aim was to study the ballistic response of the CF/PEEK and
CF/Epoxy composites subjected to a high-velocity impact by the relatively hard aluminium-
alloy projectile. As shown in Figure 6-3, two high-speed cameras were located at the side of
the target chamber and recorded at their maximum rate of 39,000 frames per second. The
kinetic energy absorbed during the impact could be determined via measuring the initial and
residual velocities of the projectile. Two pairs of infrared sensors were also used to validate the
initial velocity of the projectile as measured using the high-speed cameras.

91
CHAPTER 6

Figure 6-3. Set-up II for the high-velocity hard-projectile impact: (a) photograph and (b)
schematic.

6.2.3 Post-test inspection


The portable C-scan system were used to examine all the impacted specimens after the
soft-projectile impact experiments. A configuration designed for the 2-mm woven specimens
was operated in the C-scan system. Visual inspections, for both soft and hard impact
experiments, were performed on the rear surfaces of the specimens where the impact damage
is often found. To investigate the failure mechnisms of the specimens, the cross-section of the
mid-plane of the composites was created via a ‘Brillant 220’ cut-off machine provided by
QATM, Germany. This laser alignment aided machine, as shown in Figure 6-4, had a precision
of 1 μm and equipped with a 0.5 mm thick high concentration diamond cut-off wheel. The
wheel forwarded 0.2 mm per second and set at 2000 rpm to minimum additional damage to the
specimens.

92
CHAPTER 6

Figure 6-4. The cut-off machine equipped with high concentration diamond saw.

6.3 Experimental results


6.3.1 The soft projectile impact testing
6.3.1.1 Impact response
The impact responses of the CFRP specimens were investigated over a range of impact
velocities from 42 to 100 m.s-1 using the soft-gelatine projectiles, and were therefore studied at
different impact energy levels. Figures 6-5 and 6-6 show results for a soft-gelatine impact at
100 m.s-1 on the CF/PEEK and the CF/Epoxy specimens, respectively. Figure 6-5(a) shows the
out-of-plane displacement maps obtained from the CF/PEEK specimen under soft impact
loading, where both the loading and the unloading phases are included. The deformation profile
across the middle section of the specimen is shown in Figure 6-5(b). This CF/PEEK specimen,
which was subjected to an impact velocity of 100 m.s-1, exhibited a maximum out-of-plane
displacement of approximately 3.7 mm.

93
CHAPTER 6

(a)

(b)

Figure 6-5. Experimental results for the CF/PEEK panel impacted using a soft-gelatine
projectile with a velocity of 100 m.s-1 showing the (a) out-of-plane displacement maps and
(b) out-of-plane displacement profile during the loading and unloading phases. (Inset
picture, on right, shows a horizontal dash line where the profile section is taken.)

The out-of-plane displacement maps and deformation profiles of the CF/Epoxy


specimen, under a soft impact loading at an impact velocity of 100 m.s-1, are shown in Figures
6-6(a) and 6-6(b), respectively. In comparison to the DIC results of the CF/PEEK specimen, a
10% higher maximum out-of-plane displacement was determined for the CF/Epoxy specimen.
In addition, a slower rebound sequence for the CF/Epoxy composite was found from the
deformation profiles for the unloading phase. These differences can be explained from the
photographs of the rear surface and cross-section of the CF/Epoxy specimen shown below,
where significant impact damage was found in the central area of the specimen. The presence
of damage, which was observed to arise from matrix cracking, delamination and fibre breakage,
as discussed below, leads to a significant loss of the stiffness of the specimen. This, in turn,
leads a higher maximum out-of-plane displacement and results in a slower rebound of the
specimen in the unloading phase for the CF/Epoxy composite, compared to the CF/PEEK
composite.

94
CHAPTER 6

(a)

(b)

Figure 6-6. Experimental results for the CF/Epoxy panel impacted using a soft-gelatine
projectile with a velocity of 100 m.s-1 showing the (a) out-of-plane displacement maps and
(b) out-of-plane displacement profile during the loading and unloading phases. (Inset picture,
on right, shows a horizontal dash line where the profile section is taken.)

The 3D DIC results of the maximum out-of-plane displacement of the specimens under
four different impact velocities for both types of composites are summarised in Table 6-2. In
addition, the maximum out-of-plane displacement is plotted against the impact energy in Figure
6-7. As may be seen, the values of the maximum out-of-plane displacements increase with
increasing impact velocity. Also, at a similar impact velocity (i.e. at a similar impact energy
level) the CF/Epoxy composite undergoes a greater maximum out-of-plane displacement,
compared to the CF/PEEK composite, and this difference is most striking at the highest impact
velocity that was employed of 100 m.s-1.

95
CHAPTER 6

Table 6-2. Results of the high-velocity soft-gelatine projectile impact tests.


Impact Impact Maximum Out-of-
Specimen Projectile
Velocity* Energy* plane Displacement
Code Mass (g)
(m.s-1) (J) (mm)
CF/PEEK_1 42.4 19.9 17.9 1.6
CF/PEEK_2 58.1 19.9 33.6 2.2
CF/PEEK_3 78.1 20.0 61.0 2.9
CF/PEEK_4 100.0 19.8 99.0 3.7
CF/Epoxy_1 45.9 19.8 20.9 1.8
CF/Epoxy_2 57.5 20.0 33.1 2.3
CF/Epoxy_3 74.6 19.7 54.8 3.1
CF/Epoxy_4 100.0 19.9 99.5 4.2
*Note: Experimental error: ±2.5% in impact velocity and ±5% in impact energy.

Figure 6-7. Maximum out-of-plane displacement versus the impact energy for the CF/PEEK
and CF/Epoxy composites upon being impacted using soft-gelatine projectiles.

6.3.1.2 Visual inspections


Visual inspections were performed on the rear surfaces of the CF/PEEK and the
CF/Epoxy specimens, as shown in Figure 6-8(a) and 6-8(b), respectively, for the different
impact velocities used for the soft-gelatine projectile. Firstly, no visible damage was observed
in any of the CF/PEEK specimens. Secondly, however, matrix cracking and fibre breakage

96
CHAPTER 6

were observed, at the highest impact velocity of 100 m.s-1 on the rear surface of the impacted
CF/Epoxy composite. The cross-sections of the CF/PEEK and CF/Epoxy specimens impacted
at 100 m.s-1 are shown in Figures 6-9(a) and 6-9(b), respectively. As may be seen, again no
visible damage was observed in the CF/PEEK composites, whereas different types of damage
including matrix crack, delamination and fibre breakage were particularly observable in the
CF/Epoxy composite impacted at the highest test velocity of 100 m.s-1.

(a)

(b)

Figure 6-8. Visual inspection of the rear surface of the (a) CF/PEEK specimens
and the (b) CF/Epoxy specimens subjected to different impact velocities by the
soft-gelatine projectile. (The area represented by the red square corresponds to the
area shown on the schematic of the full specimen at the far right-hand side. The
yellow circles indicate the extent of visible damage.)

97
CHAPTER 6

(a) (b)

Figure 6-9. Photographs of the cross-section of the mid-plane of the composite: (a) CF/PEEK
and (b) CF/Epoxy specimens impacted using a soft-gelatine projectile with an impact velocity
of 100 m.s-1.

6.3.1.2 Ultrasonic C-scan maps


As noted above from Figures 6-8 and 6-9, a visual inspection of the impacted specimens
showed that significant damage was only observed in the CF/Epoxy composite tested at the
highest impact velocity, whilst no damage was ever observed in the CF/PEEK composite at
any of the impact velocities employed. The relevant C-scan images in Figure 6-10 show very
clearly an area of delamination, in both the relatively low magnification and the higher
magnification C-scan maps, for the CF/Epoxy composite impacted at the highest velocity of
100 m.s-1. (These maps show delamination as a function of the depth through the thickness of
the specimen, where the red colour is representative of the front (impacted) surface and blue
colour is representative of the rear (non-impacted) surface of the specimen.) Thus, the
ultrasonic C-scan maps reveal sub-surface damage only for the CF/Epoxy composite, and then
only significant damage when impacted at the highest velocity of 100 m.s-1. These findings are
clearly in excellent agreement with the visual inspections discussed above. Therefore, it may
again be concluded that the CF/PEEK laminates have a superior resistance to delamination
when subjected to a soft impact at 100 m.s-1.

98
CHAPTER 6

(a)

(b)
(c)

Figure 6-10. Ultrasonic C-scan maps of the (a) CF/PEEK specimens and (b) CF/Epoxy
specimens subjected to impact velocities using the soft-gelatine projectile and (c) magnified
C-scan image of the CF/Epoxy specimen impacted at the highest velocity of 100 m.s-1.

6.3.2 The hard projectile impact testing

The perforation resistance and energy absorption of both the CF/PEEK and CF/Epoxy
specimens were studied when these composites were subjected to an impact by the relatively
hard aluminium-alloy projectile over a range of impact velocities from 30 to 77 m.s-1. The
kinetic energy absorption, Ea, of the test specimen can be determined from:

1 1
Ea = mvi 2 − mvr 2 (6-1)
2 2

where m is the mass of projectile, and vi and vr are the initial impact velocity and residual
velocity (i.e. the exit or rebound velocity of the projectile), respectively. Three different impact
velocities below and above the ballistic limit, which is the lowest velocity required for the
projectile to penetrate consistently through the test specimen, were selected to investigate the
relationship between the energy absorption and the damage suffered by the two types of
composites.

The photographs obtained from the rear surface of the tested specimens are shown in
Figure 6-11. As may be seen, minor surface damage, with similar areal extents of the damage,
can be observed from both types of specimens impacted at about 30 m.s-1 (i.e. with an
associated impact energy of 12-13 J). For an impact velocity above about 50 m.s-1 (i.e. with an
99
CHAPTER 6

associated impact energy above about 40 J) the CF/PEEK specimens perforated and exhibited
damage areas which were slightly larger in size than the cross-sectional area of the projectile.
In contrast, for impact energies above about 40 J, the damage area was about three times larger
in the perforated CF/Epoxy specimens, compared to the CF/PEEK specimens. For both types
of composite there was no significant further increase in the extent of the damage area when
the impact velocity was raised to about 70 m.s-1.

(a)

(b)

Figure 6-11. Photographs of the rear surface of the (a) CF/PEEK and (b) CF/Epoxy
composites impacted using the hard aluminium-alloy projectile at various initial impact
velocities. (The visible damage area, Sd, outlined in by the broken white line, was determined
by measuring the number of pixels in the damage area compared to the total number in the
specimen of known area.)

The experimental results from using the hard aluminium-alloy projectile are
summarised in Table 6-3. (It should be noted that in Table 6-3 a positive value in the residual
velocity indicates that full perforation has occurred, whereas a negative value means that the
projectile rebounded from the front surface of the specimen. Also, the damage area, Sd, of each
specimen was determined, via a data processing programme, by measuring the number of
pixels in the damage area compared to the total number in the specimen of known area.) The
kinetic energy absorption and the damage area suffered by the composites are also plotted

100
CHAPTER 6

against the impact energy in Figure 6-12. In addition, it should be noted that no visible
deformation was observed in the aluminium-alloy projectiles post-impact.

Table 6-3. Results of the high-velocity hard aluminium-alloy projectile impact tests.
Initial Projectile Impact Residual Energy Damage
Sample
Velocity* Mass Energy* Velocity* (vr) Absorption* Area* (Sd)
Code
(vi) (m.s-1) (g) (J) (m.s-1) (Ea) (J) (cm2)
CF/PEEK_5 30.0 28.4 12.8 -14.8 9.7 0.7
CF/PEEK_6 56.7 28.2 45.3 23.2 37.7 10.0
CF/PEEK_7 76.8 28.1 82.9 50.0 47.8 11.2
CF/Epoxy_5 29.1 28.1 11.9 -15.5 8.6 0.9
CF/Epoxy_6 54.5 28.2 41.9 24.8 33.2 29.9
CF/Epoxy_7 70.7 28.3 70.7 47.8 38.4 31.3
*Note: Experimental error: ±2% in velocities, ±4% in impact energy and ±5% in damage area.

Figure 6-12. Comparison of the kinetic energy absorption (KEA) and damage area (DA)
versus the impact energy for the CF/PEEK and CF/Epoxy composites impacted using the
hard aluminium-alloy projectile at different velocities. (Note: perforation of the composites
occurred at, and beyond, an energy level of about 40 J.)

101
CHAPTER 6

In Figure 6-12 it can be seen that, as the impact energy is increased, then both the kinetic
energy absorption and the damage area increase. Once perforation occurs at an impact energy
of about 40 J, and beyond, there is, however, a far smaller increase on the energy absorption
and damage area as the impact velocity is further increased. This phenomenon has previously
been observed in the impact behaviour and failure mechanisms of composites when they are
perforated by an impacting projectile [66]. Now, the damage is considered to arise from
compressive stress waves which are generated and transmitted through the volume of the
composite material after impact. These waves are reflected when reaching the rear layer of the
composite panel to form tensile waves which now act upon the neighbouring ply interfaces and
can initiate delamination, and so lead to extensive interlaminar failure. Also, shear failures
occur when the projectile starts to penetrate the uppermost layers of the composites.
Subsequently, when the projectile reaches the delaminated layers, these layers can slide over
each other and ultimately fail completely due to tensile failure. It is for these reasons that a
larger damage area is found on the rear surface. Now, many researchers [66,67,171] have
indeed reported the phenomenon that the damage area in the composite increases relatively
rapidly with an initial increase in the impact velocity up to the ballistic limit velocity, and then
the damage area remains relatively constant in extent beyond this velocity. As was indeed
observed in the present high-velocity tests using the hard aluminium-alloy projectile.

6.4 Summary
In the present chapter, the impact performance of woven-fabric carbon-fibre
composites based upon both thermoplastic- and thermoset-matrix polymers under high-
velocity impact loading have been studied by conducting gas-gun experiments at impact
velocities of up to 100 m.s-1. The thermoplastic matrix was poly(ether-ether ketone) (PEEK)
and the thermoset matrix was a non-toughened Epoxy. The carbon-fibre reinforced-polymers
(CFRPs) were impacted using a soft (i.e. gelatine) or a hard (i.e. aluminium-alloy) projectile to
simulate either a soft bird-strike or a hard foreign-body impact (e.g. runway debris),
respectively, on composites employed in civil aircraft.

For the high-velocity soft gelatine impact tests, the 3D Digital Image Correlation (DIC)
technique was successfully employed to determine the out-of-plane displacement of the woven-
fabric composites specimens at different impact velocities, and hence different impact energy
levels. The woven-fabric CF/PEEK composites, with the same lay-up and thickness and

102
CHAPTER 6

approximately the same density as the CF/Epoxy composites, exhibited a reduced maximum
out-of-plane deformation compared to the CF/Epoxy composites. Indeed, fibre and matrix
damage on the rear of the specimen, and sub-surface damage, indicated that significant damage
was only observed in the CF/Epoxy composite, and only at the highest velocity of 100 m.s-1.

Delamination damage was indeed observed in the C-scan map of the CF/Epoxy
composite impacted at the highest velocity of 100 m.s-1. By contrast, no delamination area was
found in any C-scan maps of the CF/PEEK composite specimens. It was therefore concluded
that the CF/PEEK composites have a superior damage resistance compared to the equivalent
CF/Epoxy composites when subjected to a high-velocity impact by the soft gelatine projectile,
which simulates a bird-strike.

For the high-velocity impact tests using a hard aluminium-alloy projectile, with a hemi-
spherical head, to simulate runway debris, superior impact resistance was observed in the
CF/PEEK composites compared to the CF/Epoxy composites. The efficiency of kinetic energy
absorption decreases with the increasing impact velocity from about 30 to 75 m.s -1. The
CF/PEEK composite was more efficient in absorbing kinetic energy for these velocities and
the damage area was less extensive than for the CF/Epoxy composite. As to be expected, at
similar impact energy levels, composite specimens are more vulnerable to the harder
aluminium-alloy projectile than the soft-gelatine projectile. This is attributed to the higher
contact pressures, and their localisation, associated with the aluminium-alloy projectile.

It is important to note that the CF/PEEK composite showed improved impact


performance both when struck by hard aluminium-alloy projectiles (to simulate runway debris)
and soft gelatine projectiles (to simulate bird-strike) compared to the CF/Epoxy composite.

103
Chapter 7. Experimental study and numerical
modelling on woven-fabric thermoplastic and
thermoset laminates under high-velocity impact
loading

7.1 Introduction
This chapter presents experimental and numerical studies on the behaviour of woven-
fabric composite laminates subject to impact loading by soft projectiles to represent the impact
of a small bird or hail-stone. In this research, gas-gun experiments are performed to study
woven carbon-fibre reinforced poly (ether-ether ketone) (CF/PEEK) composites subjected to
an impact by soft-gelatine projectiles. In addition, woven carbon-fibre reinforced epoxy
(CF/Epoxy) composite specimens are also evaluated using gelatine projectiles to investigate
the effect of the matrix system on the impact response of the composites. A high-speed camera
is employed to capture the deformation of the projectiles and a three-dimensional (3D) Digital
Image Correlation (DIC) system is used to record the deformation of the impacted composite
specimens. A Finite Element (FE) model is developed to simulate the impact by a soft projectile
on the composite specimens. Good agreement is shown between the predictions from using the
FE model and the experimental results.

Note: The work presented in this chapter was previously published in the Engineering Failure Analysis
as “The behaviour of fibre-reinforced composites subjected to a soft impact-loading: An experimental
and numerical study”.

104
CHAPTER 7

7.2 Numerical modelling methods


7.2.1 Modelling the response of the projectile
The SPH approach [110] was employed to model the behaviour of the gelatine
projectile within the ‘Abaqus/Explicit 2018’ code, as discussed later. For the SPH method to
capture the response of the soft-gelatine projectile upon impact of the composite, a constitutive
law is required with suitable material properties employed for the gelatine projectile. The
model used was originally developed for ballistic impact in metals and describes an isotropic
elastic-plastic material subjected to relatively low pressures with an equation of state (EOS)
describing the hydrodynamic pressure versus volume behaviour at high pressures. The linear
Mie-Grüneisen EOS was employed to define the coupled equations for pressure and internal
energy [148]. The most common form for this EOS is given by:

𝑝 − 𝑝𝐻 = 𝛤𝜌(𝐸𝑚 − 𝐸𝐻 ) (7-1)

where 𝑝 is the pressure which is defined as positive in compression. The Hugoniot pressure,
𝑝𝐻 , is a function only of the density and can be ascertained from fitting to the experimental
data. The parameters 𝐸𝑚 and 𝐸𝐻 are the internal energy per unit mass and the specific energy
per unit mass (i.e. the Hugoniot energy), respectively. The parameter, 𝜌, is the current density
of the gelatine projectile. The parameter, 𝛤, is the Grüneisen ratio and is defined by:

𝜌0
𝛤 = 𝛤0 (7-2)
𝜌

where 𝛤0 is a material constant and 𝜌0 is the reference density of the gelatine projectile. The
specific energy per unit mass, 𝐸𝐻 , is related to the Hugoniot pressure by:

𝑝𝐻 𝜂
𝐸𝐻 = (7-3)
2𝜌0

105
CHAPTER 7

where 𝜂 = 1 − 𝜌0 /𝜌 and 𝜂 is the nominal volumetric compressive strain. The elimination of


𝛤 and 𝐸𝐻 from the above equations yields:

𝛤0 𝜂
𝑝 = 𝑝𝐻 (1 − ) + 𝛤0 𝜌0 𝐸𝑚 (7-4)
2

In the above equation, the pressure, 𝑝, is a function of the Hugoniot pressure, 𝑝𝐻 , and
the nominal volumetric compressive strain, 𝜂 . Once the relationship between 𝑝𝐻 and 𝜂 is
defined, the pressure, 𝑝, can be expressed as a single-variable function. To achieve this, the
linear 𝑈𝑠 versus 𝑈𝑝 relationship was employed to fit the curve of the Hugoniot pressure versus
the nominal volumetric compressive strain. The term 𝑈𝑠 is the shock-wave velocity. The term
𝑈𝑝 is the particle velocity of the projectile and the measured value of 𝑈𝑝 was assigned to all
the 8-node linear-brick (C3D8R) elements when the FE model was started, as discussed in
detail later. However, immediately after initial contact of the projectile with the composite,
these elements for the projectile were converted to continuum particle (PC3D) elements and
the value of 𝑈𝑝 assigned to the particle elements was then continually updated based upon the
loading conditions on the particles after the initial contact. Assuming the usual linear 𝑈𝑠 versus
𝑈𝑝 relationship, then the Hugoniot pressure versus the nominal volumetric compressive strain
equation is given by:

𝜌0 𝑐02 𝜂
𝑝𝐻 = (7-5)
(1 − 𝑠𝜂)2

where the fitting coefficient, 𝑠, is the slope of the linear relationship between 𝑈𝑠 and 𝑈𝑝 :

𝑈𝑠 = 𝑐0 + 𝑠𝑈𝑝 (7-6)

where 𝑐0 is the reference speed of sound in the gelatine projectile. With the above assumptions,
the relationship between the pressure, 𝑝, and the nominal volumetric compressive strain, 𝜂 may
now be written as:

106
CHAPTER 7

𝜌0 𝑐02 𝜂 𝛤0 𝜂
𝑝= (1 − ) + 𝛤0 𝜌0 𝐸𝑚 (7-7)
(1 − 𝑠𝜂)2 2

Thus, in the FE model, see below, Equations (7-5) and (7-6) were employed to define
the parameters in the EOS for modelling the gelatine projectile and Equation (7-7) was used to
predict the contact pressure between the gelatine projectile and the composite.

7.2.2 Modelling the response of the composites


7.2.2.1 The intralaminar damage model
The model for predicting the initiation of any intralaminar damage was implemented
within the ‘Abaqus/Explicit 2018’ FE code, and was based upon Hashin’s theory [148,151,172],
which assumes that the undamaged composite exhibits linear-elastic behaviour. In Hashin’s
damage model, four different types of damage mechanisms, which arise from tensile fibre
failure, compressive fibre failure, tensile matrix failure and compressive matrix failure, are
employed to capture the initiation of intralaminar damage in the unidirectional-fibre sub-plies.
Corresponding to the damage initiation mechanisms defined in Hashin’s criteria, four damage
parameters were implemented in the damage evolution model, which arise from four failure
modes mentioned above. Three damage variables which reflect fibre damage, matrix damage
and shear damage can then be derived from these damage parameters. When any of these three
damage variables meet the stated condition, then the these damaged elements are deleted from
the model [148]. The related equations and derivation are detailed in Chapter 5.2.2.

7.2.2.2 The interlaminar damage model


Interlaminar damage typically involves the initiation and growth of delamination
between the plies that make up the composite laminate and this was captured using the Abaqus
built-in surface-based cohesive (i.e. interface) element using an energy-release approach [148].
The interface element was described via a cohesive (i.e. damage) surface law [118,131,154–
156] and is in the form of a bilinear cohesive law for a linear-softening material model. This
damage law is divided into two steps. Before the initiation of any delamination, the relationship
possesses linear-elastic behaviour. Once the damage criterion is satisfied, the cohesive stiffness
degrades linearly until separation of the interface, i.e. the delamination now propagates when

107
CHAPTER 7

the maximum failure displacement is attained. Finally, the initiation and growth of any
intralaminar damage, as discussed above, significantly influences the extent of interlaminar
damage and hence these two damage modes were modelled to be interactive in the Abaqus
simulation, see Figure 5-1 in Chapter 5. The related equations and derivation are detailed in
Chapter 5.2.3.

7.3 Experimental test specification and methods


7.3.1 Projectiles and composites
A 6 wt.% gelatine projectile, as the soft impator to simulate birdstrike, was employed
in this high-velocity soft impact tests. The detailed procedure, to manufacture the gelatine
projectiles, is presented in Table 6-1 in Chapter 6. The gelatine projectiles had a nominal weight
of 20 g, a nominal diameter of 24 mm and a nominal length of 40 mm. Due to their relatively
low hardness, the gelatine projectiles initially tended to deform during the launching event from
the gas-gun. To eliminate this problem, a plastic sabot was developed to maintain the shape of
the gelatine projectile during the acceleration phase of the impact tests.

A woven T300 carbon-fibre reinforced PEEK composite and a woven T300 carbon-
fibre reinforced ‘Toray 3631’ epoxy composite (as detailed in Chapter 3.2.2) were studied. The
woven carbon-fibre ply possessed a [0˚-90˚] architecture and had a nominal thickness of 2 mm.
For the DIC measurement, the specimens were first painted on the rear-face using a white matt
paint and then ‘speckled’ using a paintbrush to form the matt-black pattern.

7.3.2 Experimental investigations


The gas-gun (as introduced in Chapter 3.4.2) was employed to accelerate the projectiles
in the impact tests. The same steel fixture with an opening of 70 × 70 mm, used in the study
illustrated in Chapter 6, was employed to support the specimens. A new projectile and a new
sabot were employed for each impact test. A 3D DIC system was used to measure the
deformation of the rear-face of the specimens during impact loading. Two ‘Phantom Miro
M/R/LC310’ high-speed cameras, supplied by Vision Research Phantom, USA, were
employed. A pair of identical ‘Nikon’ lenses, with a fixed focal length of 50 mm, supplied by
Nikon, UK, were used with these two cameras. During the tests, the recording rate of these two

108
CHAPTER 7

cameras was set at 39,000 frames per second and they were triggered simultaneously by the
signal generated from the infrared sensors.

After the impact experiments, visual inspections were undertaken on the composite
specimens and photographs were taken from the rear-faces of the post-impacted specimens. In
general, the type of damage suffered by the composites on the rear-face could be categorised
as: (a) no visible damage present, (b) cracking observed, (c) fracture having occurred, and (d)
perforation (i.e. penetration of the projectile through the specimen) having occurred. The main
difference between ‘type (b) cracking’ and ‘type (c) fracture’ is whether there was fibre
breakage observed. For ‘type (b) cracking’ this was defined as when cracks were only observed
in the matrix. However, for ‘type (c) fracture’, fibre failure was also observed. Schematics of
these descriptions for status of the post-impacted composites are shown in Figure 7-1.

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

Figure 7-1. Schematics of the types of post-impact damage on the rear-face of the
composites: (a) no visible damage, (b) cracking, (c) fracture and (d) perforation.

7.4 Experimental results


7.4.1 Deformation of the gelatine projectile
Figure 7-2 shows the deformation of the gelatine projectile recorded by a high-speed
camera during an impact with the CF/PEEK composite specimen for an impact energy of 37 J.
Within the resolvable time intervals, the time, t, corresponding to the initial contact was defined
as 0.0 ms, as shown in Figure 7-2(b). It was found that at the beginning of the impact event (i.e.
t = 0.0 ms) the shape of the gelatine projectile was well preserved, which ensured that the
gelatine projectile impacted the centre of the specimen and then deformed symmetrically.
However, in Figure 7-2(f) for t = 0.4 ms, the gelatine projectile can clearly be seen to be flowing
freely after impact.

109
CHAPTER 7

-0.1 ms 0 ms 0.1 ms 0.2 ms 0.3 ms 0.4 ms

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

Figure 7-2. Deformation history of the gelatine projectile for a 37 J impact energy impacting
the CF/PEEK composite.

7.4.2 Effects of the impact energy of the gelatine projectile


To study the effects of the impact energy on the response of the CF/PEEK composites
subjected to soft impact-loading, these composites were impacted using gelatine projectiles
fired at four different impact velocities, and hence with different impact energies. The testing
configurations for investigating the effects of the impact energy on the impact response of the
CF/PEEK specimens are given in Table 7-1.

Table 7-1. Test configurations for investing the effects of the impact velocity and energy on
the CF/PEEK composites.
Test Projectile Projectile mass (g) Impact velocity (m.s-1) Impact energy (J)
GCP-I Gelatine 20 ± 0.5 61 ± 2.5% 37 ± 5%
GCP-II Gelatine 19 ± 0.5 75 ± 2.5% 53 ± 5%
GCP-III Gelatine 20 ± 0.5 80 ± 2.5% 64 ± 5%
GCP-IV Gelatine 20 ± 0.5 85 ± 2.5% 72 ± 5%

7.4.2.1 Comparison of the Digital Image Correlation (DIC) results


The 3D DIC system was employed to measure the major strain and out-of-plane
displacement on the rear-faces of the composites. The main DIC results obtained from the
CF/PEEK composites, impacted by gelatine projectiles at different energy levels, are
summarised in Table 5. (It should be noted that due to fracture of the rear-face during ‘Test

110
CHAPTER 7

GCP-IV’ when an energy level of 72 J was used, no accurate value for the maximum major
strain could be obtained from the DIC results for this test.) As the impact energy is steadily
increased, the maximum major strain and maximum OOP displacement both increased in value.

Table 7-2. Main DIC results for the CF/PEEK composites impacted by the gelatine projectiles.
Impact velocity Impact energy Maximum OOP
Test
(m.s-1) (J) displacement (mm)
GCP-I 61 ± 2.5% 37 ± 5% 3.9 ± 3%
GCP-II 75 ± 2.5% 53 ± 5% 4.2 ± 3%
GCP-III 80 ± 2.5% 64 ± 5% 4.6 ± 3%
GCP-IV 85 ± 2.5% 72 ± 5% 4.8 ± 3%

Figures 7-3 present the typical DIC results obtained from the CF/PEEK composite
impacted by a gelatine projectile with impact energy of 37 J. The OOP displacement contours,
corresponding to different times during the impact tests, were also obtained from the DIC
results and are shown in Figure 7-3(a) for an impact energy of 37 J. Similarly, the OOP
displacements along the horizontal mid-section, during the loading and unloading process of
the specimen, were also obtained and are shown in Figure 7-3(b).

111
CHAPTER 7

(a)

(b)

Figure 7-3. CF/PEEK composites impacted at a 37 J energy level: (a) the OOP
displacement contours and (b) the evolution of the OOP displacement profiles (in intervals
of 0.025 ms) during loading and unloading. (Inset picture, on right, shows a horizontal
solid line where the profile section is taken.)

7.4.2.2 Comparison of the Digital Image Correlation (DIC) results


Representative photographs taken of the rear-faces of two of the gelatine-impacted
CF/PEEK specimens are shown in Figure 7-4, along with corresponding magnified images of
the central area. In Figure 7-4(a), where the CF/PEEK composite was impacted by a gelatine
projectile with energy of 37 J, no visible damage was observed. The same observation, of no
visible damage, was recorded for the CF/PEEK tests conducted at impact energy levels of 53 J
and 64 J. In contrast, the CF/PEEK composite impacted using a gelatine projectile with an
impact energy of 72 J has suffered ‘type (c)’ fracture damage, with cracking in the matrix
mainly being confined to the central area of the specimen, as shown in Figure 7-4(b). Further,
obvious fibre breakage was observed in this CF/PEEK composite specimen. Thus, it can be
concluded that there is a critical impact energy between about 64 J and 72 J at which visible
damage in the CF/PEEK composite is initiated.

112
CHAPTER 7

(a)

(b)

Figure 7-4. Photographs of the rear-faces of the CF/PEEK composites after impact: (a) for
an energy of 37 J (‘Test GCP-I’) and (b) for an energy of 72 J (‘Test GCP-IV’).

7.4.3 Effects of the matrix system


To study the effects of the employed matrix system on the impact response of the
composite laminates, CF/epoxy composite specimens were also impacted, using a soft-gelatine
projectile, at an energy level of 38 J. The details of the testing conditions are summarised in
Table 7-3 and the results are shown in Table 7-4 and Figure 7-5. As may be seen, the main
effect of the matrix selected for the carbon-fibre composite is that the CF/PEEK composite
(‘Test GCP-I’) impacted at an energy level of 37 J did not show any visible damage, whilst the
CF/epoxy composite (‘Test GCE-I’) showed significant damage with ‘type (b) cracking’ being
recorded.

113
CHAPTER 7

Table 7-3. Gas-gun test conditions to study the effect of the matrix system.
Matrix Projectile Impact velocity Impact
Test Projectile
system mass (g) (m.s-1) energy (J)
GCP-I Gelatine PEEK 20 ± 0.5 61 ± 2.5% 37 ± 5%
GCE-I Gelatine Epoxy 20 ± 0.5 62 ± 2.5% 38 ± 5%

Table 7-4. Results from the CF/PEEK and CF/Epoxy composites impacted by the gelatine
projectiles.
Matrix Impact velocity Energy Maximum Maximum OOP
Test
system (m.s-1) (J) major strain displacement (mm)
GCP-I PEEK 61 ± 2.5% 37 ± 5% 0.013 ± 3% 3.9 ± 3%
GCE-I Epoxy 62 ± 2.5% 38 ± 5% N/A 4.0 ± 3%

(a)

(b)

Figure 7-5. The rear-faces of the specimens after impact: (a) the CF/PEEK composite
impacted at a 37 J energy level and (b) the CF/epoxy composite impacted at a 38 J energy
level.

114
CHAPTER 7

7.5 The Finite Element (FE) model


7.5.1 Model definition
As a discussed earlier, in order to model the soft-body impact on the composite test
specimens a Finite-Element (FE) model was developed based upon a commercial software
code, ‘Abaqus/Explicit 2018’. Within the FE model, the gelatine projectile was modelled using
the Smoothed Particle Hydrodynamics (SPH) modelling technique [110]. The SPH method is
a meshless Lagrangian technique where the solid FE mesh for the gelatine impactor is replaced
by a set of discrete interacting particles. The gelatine projectile was first modelled using 8-node
linear-brick (C3D8R) elements. However, upon initial contact of the projectile with the
composite target specimen, these elements were converted to continuum particle (PC3D)
elements, see Figure 7-6. The characteristic length for the PC3D elements was 0.5 mm, which
was equivalent to half of the element size that was used for modelling the gelatine projectile
with the C3D8R elements. The total mass of the projectile was equally distributed between all
the 8-node linear-brick (C3D8R) elements or all the continuum particle (PC3D) elements.

Figure 7-6. The FE model with PC3D elements.

Turning to the modelling of the composite specimen, the damage theories discussed
earlier [148,151,172] were originally developed for unidirectional fibre-reinforced composite

115
CHAPTER 7

plies. Hence, the [0˚-90˚] woven carbon-fibre ply used for the CF/PEEK and CF/epoxy
composites was represented as two unidirectional-fibre sub-plies, joined at right angles to the
fibre direction. Thus, in the FE modelling, two unidirectional-fibre sub-plies were first created,
with the thickness of each of the unidirectional-fibre sub-plies (i.e. 0.125 mm) being equal to
half that of the thickness of the equivalent [0˚-90˚] woven-fibre composite ply (i.e. 0.25 mm).
These two unidirectional-fibre sub-plies were placed at right angles and then joined using ‘tie
constraints’, to form a single equivalent [0˚-90˚] woven-fibre composite ply, which has the
same in-plane properties as the actual woven-fibre composite ply that was used in the
composite specimens, see Figure 7-7. The elements employed in the FE model for the
composite target test specimens were 8-node quadrilateral in-plane general-purpose continuum
shell (SC8R) elements, with an element size of 1 mm × 1 mm. The interfaces between the
composite plies were modelled using the cohesive surface law, which is again a built-in sub-
routine within the ‘Abaqus/Explicit 2018’ code [135,149,173]. The boundary conditions
employed in the model were the same as those used in the gas-gun experiments. A general
contact algorithm was used to govern the global contact in the numerical modelling and a
friction coefficient of 0.2 was adopted for the global contact [174–176].

Figure 7-7. The creation of a single equivalent [0o-90o] woven-fibre reinforced composite
ply.

7.5.2 Input parameters


In order to use the SPH method for capturing the response of the soft-gelatine projectile,
an equation of state (EOS) with suitable input parameters, as shown in Equations (7-5) and (7-
6), is required for the modelling of the gelatine projectiles, see Chapter 7.2.1. The input
116
CHAPTER 7

parameters required for the numerical modelling of the gelatine projectiles are shown in Table
7-5. For the composite specimen, it was defined using continuum shell elements and only the
in-plane material properties are then required for the numerical modelling. However, the values
of the cohesive stiffness, maximum cohesive strength and the various fracture energies do need
to be inputted into the sub-routine which simulates the damage evolution in the composite via
a linear-softening material model embedded in a bilinear cohesive law. The relevant material
properties of the CF/PEEK and CF/Epoxy composites required for the FE modelling studies
may be found from the literature [133,147,153,160–162,177,178] and are given in Table 7-6.

Table 7-5. Input properties for the FE modelling of the soft-gelatine projectile [37-39].
Reference Dynamic Reference speed Slope of the
Grüneisen
Properties density viscosity of sound Us versus Up
ratio
(g.cm-3) (MPa.s) (mm.s-1) curve

Values 1.06 1 × 10-6 𝑐0 = 1.45 × 106 𝑠 = 1.87 𝛤 = 1.09

117
CHAPTER 7

Table 7-6. Input properties for the FE modelling studies of the composite [133,141,147,153,160–163].
Property Unidirectional CF/PEEK sub-ply Unidirectional CF/epoxy sub-ply
Moduli (𝐆𝐏𝐚) 𝐸11 = 127; 𝐸22 = 10.3; 𝐺12 = 5.7 𝐸11 = 125; 𝐸22 = 8.7; 𝐺12 = 4.3
Poisson`s ratio 𝜈12 = 0.3 𝜈12 = 0.3

𝑋 𝑇 = 2070; 𝑌 𝑇 = 85 𝑋 𝑇 = 1930; 𝑌 𝑇 = 41
Strength values (𝐌𝐏𝐚) 𝑋 𝐶 = 1360; 𝑌 𝐶 = 276 𝑋 𝐶 = 1250; 𝑌 𝐶 = 254
𝑆 𝐿 = 𝑆 𝑇 = 186; 𝑆 𝐿 = 𝑆 𝑇 = 110

𝐺𝐼𝑐 |𝑓𝑡 = 218; 𝐺𝐼𝑐 |𝑓𝑐 = 104 𝐺𝐼𝑐 |𝑓𝑡 = 201; 𝐺𝐼𝑐 |𝑓𝑐 = 92;
Ply fracture energies (𝐤𝐉/𝐦𝟐 )
𝐺𝐼𝑐 |𝑚𝑡 = 1.7; 𝐺𝐼𝑐 |𝑚𝑐 = 2.0 𝐺𝐼𝑐 |𝑚𝑡 = 0.6; 𝐺𝐼𝑐 |𝑚𝑐 = 1.5

Interlaminar fracture energies+ (𝐤𝐉/𝐦𝟐 ) 𝐺𝐼𝑐 = 1.7 ; 𝐺𝐼𝐼𝑐 = 2.0 𝐺𝐼𝑐 = 0.6 ; 𝐺𝐼𝐼𝑐 = 1.5
Benzeggagh–Kenane mode-mix exponent+ 𝜂𝐵𝐾 = 1.09 𝜂𝐵𝐾 = 2.09
0 0 0 0 0 0
Initial cohesive law strength (𝐌𝐏𝐚) 𝑡33 = 43; 𝑡31 = 𝑡32 = 50 𝑡33 = 20; 𝑡31 = 𝑡32 = 34
Initial cohesive law stiffness (𝐌𝐏𝐚/𝐦𝐦) 𝑘 = 6.4 × 105 𝑘 = 6.2 × 105

Note: +These properties are for interlaminar failure between two of the [0o-90o] woven-fibre plies.
𝐺𝐼𝑐 and 𝐺𝐼𝐼𝑐 are the Mode I and Mode II interlaminar fracture energies between two [0o-90o] woven-fibre composite plies.
𝐺𝐼𝑐 |𝑓𝑡 and 𝐺𝐼𝑐 |𝑓𝑐 are the tensile and compressive ply fracture energies of the unidirectional-fibre sub-plies in the longitudinal fibre-direction.
𝐺𝐼𝑐 |𝑚𝑡 and 𝐺𝐼𝑐 |𝑚𝑐 are the tensile and compressive ply fracture energies of the unidirectional-fibre sub-plies in the transverse to fibre direction.

118
CHAPTER 7

7.5.3 Implementation of the model


In the computation process a computation time-step was performed for every
appropriate single element in the FE model. The numerical model is stopped when the defined
total time for the impact event has expired. The flow chart of the main FE model is as shown
in Figure 5-1 in Chapter 5.

7.6 Model validation and application


7.6.1 Validation of the model
7.6.1.1 The deformation of the gelatine projectile
The deformation histories of the gelatine projectile obtained from the experimental
studies and predicted using the FE model for an impact test conducted at an energy level of 37
J on the CF/PEEK composite (i.e. ‘Test GCP-I’) are compared in Figure 7-8. The experimental
results show that, after the initial contact with the composite specimen, the front of the gelatine
projectile started to deform and flow to the periphery of the composite specimen.
Correspondingly, the modelling results show a similar phenomenon, as shown in Figure 7-8(b).
At a later stage of the impact event, see Figure 7-8(e), most of the gelatine projectile has
deformed, flowed and spread over the surface of the composite specimen, and again the
modelling studies accurately capture this behaviour of the gelatine projectile. Thus, the
comparison between the experimental and numerical modelling results reveal that the SPH
model for the relatively soft-gelatine projectile can indeed reproduce the experimental
behaviour of the soft-gelatine projectile used in the gas-gun impact experiments.

119
CHAPTER 7

0 ms 0.1 ms 0.2 ms 0.3 ms 0.4 ms


Experiment
Simulation

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


Figure 7-8. Deformation of the gelatine projectile obtained from the experimental studies and as
predicted from the numerical FE model for the CF/PEEK composites at an impact energy of 37 J.

7.6.1.2 The CF/PEEK composites


Based on the DIC results obtained from the experiment conducted at an impact energy
of 37 J using the gelatine projectile (i.e. ‘Test GCP-I’), the major strain and out-of-plane (OOP)
displacement histories of the centre point for the rear-face of the composite test specimen can
be extracted. The values of the maximum major strain and central OOP displacement predicted
from the FE modelling studies are compared with the corresponding experimental results in
Figure 7-9(a), and good agreement may be seen. To further confirm the accuracy of the
numerical FE model, the predicted central OOP displacement versus time trace was also
compared with the corresponding experimental results, see Figure 7-9(b). It can be seen from
these results that, although the modelling studies gave somewhat lower maximum values than
the experimentally measured values, the general trend and overall response of the composites
were predicted extremely well using the numerical FE model. The slightly lower prediction
values may be due to curvature effects in the woven material which the model could not fully
capture.

120
CHAPTER 7

(a)

(b)

Figure 7-9. Comparison between the predicted and experimental results for the CF/PEEK
composite at an impact energy of 37 J: (a) the maximum major strain and the out-of-plane
(OOP) displacement and (b) the central OOP displacement versus time trace.

The next step is to assess the capability of the numerical FE model that has been
developed to predict the impact damage created in the composite by the impact event, and two
impact energies levels of 37 J and 72 J were so modelled. The experimental and predicted
extents of damage at these two energy levels, which resulted in the CF/PEEK composites, are
shown in Figures 7-10(a) and 7-10(b), respectively. (The ‘DAMAGESHR’ shown in the legend
corresponds to the shear damage.) It was found that, at an energy level of 37 J, the prediction
from the FE numerical modelling studies was that no visible impact damage would have been
suffered by the composite specimen. This finding is in excellent agreement with the
experimental results. When an impact energy level of 72 J was modelled, failure was predicted
to be present only in the central area of the CF/PEEK composite, as shown in Figure 7-10(b).
The experimental results revealed that some damage had indeed occurred in this region of the
composite. In addition, the extent of the damage, as determined from the post-impact

121
CHAPTER 7

experimental observations on this composite specimen, is accurately predicted by the


numerical studies.

(a)

(b)

Figure 7-10. The experimentally measured and predicted degrees of damage resulting in the
rear-face of the CF/PEEK composites at impact energies of (a) 37 J and (b) 72 J.

7.6.2 Application of the model


7.6.2.1 Predicting the deformation of the CF/PEEK and CF/Epoxy
composites
To model the effects of the matrix system on the impact response of the composites,
the central OOP displacement was predicted from the FE model for the CF/PEEK specimens
impacted at a 37 J energy level and the CF/Epoxy specimens impacted at a 38 J energy level,
as shown in Figure 7-11(a). The central OOP displacement versus time traces predicted for the
CF/PEEK composite and for the CF/epoxy composite exhibited a very similar behaviour up to
a peak value of the displacement followed by a gradual decrease. Figure 7-11(b) shows a
comparison of the maximum central OOP displacements predicted in the FE model for the
CF/PEEK (‘Test GCP-I’ at 37 J) and the CF/epoxy (‘Test GCE-I’ at 38 J) composites. When

122
CHAPTER 7

impacted, the CF/epoxy composite (‘Test GCE-I’) is predicted from the FE modelling to
undergo a maximum central OOP displacement of 3.9 mm, which is marginally higher than
that of 3.7 mm for the CF/PEEK composite (‘Test GCP-I’). These predicted values of the
central OOP displacement for the two types of composite are also compared with the
experimental results in Table 7-7, where very good agreement may be seen between the
experimental measurements and the FE modelling simulations. The OOP displacement
response for CF/PEEK and CF/Epoxy are very similar as both composites have the same
carbon fibres with similar volume fraction.

(a)

(b)

Figure 7-11. Predicted (a) central out-of-plane (OOP) displacement versus time trace and (b)
the experimentally measured and predicted maximum OOP displacements for the CF/PEEK
impacted at 37 J and the CF/epoxy impacted at 38 J.

123
CHAPTER 7

Table 7-7. Comparison of the experimentally measured and the numerically predicted
maximum central out-of-plane (OOP) displacement.
𝒅𝒎 −𝒅𝒆
Composite Energy level Experiment (𝒅𝒆 ) Simulation (𝒅𝒎 ) Deviation (| 𝒅𝒆
| × 100%)

CF/PEEK 37 J 3.9 mm 3.7 mm 5.1%


CF/Epoxy 38 J 4.0 mm 3.9 mm 2.5%

7.6.2.2 Predicting the post-impact damage of the composites


A comparison of the post-impact damage in the composites obtained from the
experiments and the FE numerical modelling results for the CF/PEEK and the CF/epoxy
composites is shown in Figures 7-12(a) and 7-12(b), respectively. It can be seen that the
predicted results for the CF/PEEK did not show any damage, which agrees fully with the
experimental observations. On the other hand, the modelling results for the CF/epoxy predicted
that some centrally-located damage would occur, which was indeed observed in the
experimental studies. The evolution of damage was observed by plotting the derived damage
variable, 𝑑𝑠 , which is dependent on fibre and matrix failure, 𝑑𝑓 and 𝑑𝑚 respectively, as
defined by Equation (5-12) in Chapter 5. The damage observed in CF/epoxy on the rear face
was mostly localised matrix and fibre failure at the centre of the panel. The amount of energy
expended in damage of the composite sample for CF/PEEK and CF/epoxy was very small
relative to the incident impact energy. Most of the incident impact energy is transformed into
elastic energy in the specimen which is then dissipated in friction at the support fixtures and in
intrinsic damping, as the specimen vibrates after impact. Certainly, some of the incident impact
energy is dissipated in plastic flow of the projectile.

124
CHAPTER 7

(a)

(b)

Figure 7-12. Comparison of the damage obtained from the experiments and
the FE modelling: (a) the CF/PEEK composite impacted at 37 J and (b)
CF/epoxy composite impacted at 38 J.

7.6.2.3 The deformation of the gelatine projectile


The numerical FE model was also employed to predict the average contact pressure
between the soft-gelatine projectile and the composite specimen by using Equation (7-7), see
Chapter 7.2.1. This parameter could not be readily experimentally measured in the gas-gun
experiments. The contact pressure versus time histories were obtained from the FE models for
(a) the CF/PEEK composite impacted at an impact energy of 37 J and (b) the CF/epoxy
impacted an impact energy of 38 J, and the results are shown in Figure 7-13(a). It can be seen
that the CF/PEEK and the CF/epoxy composites suffered a very similar average contact
pressure history, with an initial short duration compressive phase giving rise to a relatively high
initial contact pressure. The predicted maximum average contact pressures for the CF/PEEK
and the CF/epoxy impact tests, when the relatively soft-gelatine projectile was used, are 10.7
MPa and 9.8 MPa, respectively, as shown in Figure 7-13(b).

125
CHAPTER 7

(a)

(b)

Figure 7-13. Numerical predictions from the FE model for: (a) the average contact pressure
versus time history and (b) the maximum average contact pressure. For the CF/PEEK
composite impacted at an energy of 37 J and (b) the CF/epoxy impacted an energy of 38 J.

7.7 Summary
This chapter has focussed on experimental and numerical studies of the response of
woven-fabric CFRPs under impact loading by a soft projectile. A simple but reliable technique
was proposed for the preparation of the relatively soft-gelatine projectiles and a plastic sabot
was employed to maintain the shape of the gelatine projectile upon being launched from the
gas-gun. A high-speed camera was used to record the deformation of the projectile during the
impact event. The recorded frames showed that the gelatine projectile behaved as a
viscoelastic-plastic fluid. The gas-gun tests were firstly performed using woven carbon-fibre
reinforced poly(ether-ether ketone) (CF/PEEK) composite specimens, using the gelatine
projectiles, at four different impact energy levels. Secondly, to investigate the effects of the
matrix system on the impact response of the composites, woven carbon-fibre reinforced epoxy
126
CHAPTER 7

(CF/Epoxy) were impacted using the gelatine projectiles. The experimental results
demonstrated that the CF/epoxy composite exhibited a lower impact resistance and suffered
more impact damage, compared with the CF/PEEK composite, when struck by the gelatine
projectiles using a similar impact energy.

A finite-element (FE) numerical model was developed, which was based on the
‘Abaqus/Explicit 2018’ commercially-available software code, for predicting the behaviour of
the projectile and the composite test specimen during the impact event. The FE numerical
model has enabled (a) the deformation, (b) the initiation of damage, and (c) the evolution of
such damage in the composite target specimens to be predicted, as well as the deformation and
flow behaviour (and the contact pressure) of the projectile. The results from the numerical
studies have been found to be in very good agreement with the experimental results.

In terms of design, woven architectures are often employed on the outside of composite
laminates to generate a hybrid architecture. Woven composites do not have the stiffness of an
equivalent laminate carbon-fibre material but they have the advantage that delamination and
interfacial cracking does not occur so readily as has been shown by the above modelling and
experiments. CF/PEEK is a very attractive material as the threshold for damage is higher than
an equivalent CF/epoxy and this has also been confirmed by the modelling and experiments.

127
Chapter 8. Conclusions

8.1 Introduction
In this Ph.D. project, the performance of unidirectional cross-ply and woven-fabric
carbon fibre composites, based upon thermoplastic-poly (ether-ether-ketone) (PEEK) and
thermoset-epoxy polymers, under low-velocity and high-velocity impact tests using a rigid (i.e.
steel or aluminium-alloy) impactor or a soft (i.e. gelatine) impactor, were studied. The three-
dimensional (3D) Digital Image Correlation (DIC) technique was successfully employed to
determine the out-of-plane (OOP) displacement of the composite specimens at different impact
velocities, and hence different impact energy levels. The damage inflicted in the composite
specimens was assessed by both a visual inspection and by ultrasonic C-scanning. An elastic
finite-element (FE) model was developed to simulate the experimental results obtained from
both the low-velocity and high-velocity impact tests. The results from this model were
quantitatively validated against the impact response of the composites and captures the
essential aspects of their impact behaviour and can be used to optimise the impact performance
of such materials in industrial applications. The outcomes of the studies are summarised as
follows.

8.2 Effects of the matrix system employed on the


performance of CFRPs
The impact behaviour of a thermoplastic CF/PEEK composite and a toughened-
thermoset CF/Epoxy composite, which possessed the same cross-ply lay-up of [03/903]2s, under
low-velocity and high-velocity impact using a rigid hemispherical-shaped impactor has been
investigated. For the low-velocity drop-weight tests at the impact energies of 4.5, 7.5 and 10.5
J, the experimental results demonstrated that, at the same impact energy, the CF/PEEK

128
CHAPTER 8

composite always possessed a superior impact performance compared with the CF/Epoxy. This
was reflected by the CF/PEEK composite requiring a higher load to initiate damage and a
smaller area of delamination damage being present in the CF/PEEK composite post-impact. In
the high-velocity tests at an impact energy of 10.5±0.3 J, the results revealed that the CF/PEEK
and CF/Epoxy composites exhibited a similar damage pattern. However, the CF/Epoxy
composite clearly exhibits the higher average damage area but with a significantly larger degree
of scatter. The same conclusion can be found in the high-velocity impact tests on the woven-
fabric composites using a soft gelatine projectile or a hard aluminium-alloy projectile with a
hemi-spherical head. In the soft gelatine impact tests, the CF/PEEK composites exhibited a
reduced maximum OOP deformation and less delamination damage compared to the CF/Epoxy
composites. It was therefore concluded that the CF/PEEK composites have a superior damage
resistance compared to the equivalent CF/Epoxy composites. In the hard impact tests, the
CF/PEEK composite was more efficient in absorbing kinetic energy for these velocities and
the damage area was less extensive than for the CF/Epoxy composite. In a word, CF/PEEK is
a very attractive material as the threshold for damage is higher than an equivalent CF/epoxy
and this has also been confirmed by the modelling and experiments.

8.3 Effects of the loading rate on the impact performance


of CFRPs
As measured from the C-scans, the damage areas for the unidirectional cross-ply
CF/PEEK and CF/Epoxy composite specimens from both the low-velocity and the high-
velocity tests but with a similar impact energy of approximately 10.5 J, has been compared.
Both the CF/PEEK and CF/Epoxy composite specimens clearly exhibit very significantly
larger damage areas when subjected to the high-velocity tests of approximately 54.4±1.0 m.s-
1
, when compared with the damage areas associated with the low-velocity tests at 2.56 m.s-1,
by a factor of about four for both types of composite. This observation undoubtedly arises from
several causes, since in the high-velocity tests (a) the CF/PEEK and CF/Epoxy composite
specimens have less time to absorb the impact energy and therefore the deformation in the
specimen is initially more localised around the impact site with an increased curvature of the
specimen local to the impact site, which will induce more extensive localised delaminations
throughout the composite plies, (b) these extensive localised delaminations will then readily
propagate along the various 0o/90o ply interfaces and (c) the values of the interlaminar and

129
CHAPTER 8

matrix fracture energies will be somewhat lower for both the CF/PEEK and CF/Epoxy
composites due to the relatively high strain-rate effects associated with the high-velocity test
and therefore delaminations and matrix damage will initiate and evolve more readily in such
tests. The overall consequence is that the damage inflicted is more extensive in the high-
velocity tests compared with the low-velocity tests, as indeed is observed from the greater area
of delaminations in the C-scans of the high-velocity tests.

8.4 Effects of the hardness of the projectile on the impact


response of CFRPs
The impact performance of woven-fabric CF/PEEK and CF/Epoxy under high-velocity
impact using a soft gelatine or a hard aluminium-alloy projectile to simulate either a soft bird-
strike or a hard foreign-body impact (e.g. runway debris), respectively. A high-speed camera
was used to record the deformation of the projectile during the impact event. The recorded
frames showed that the gelatine projectile behaved as a viscoelastic-plastic fluid. As to be
expected, at similar impact energy levels, composite specimens are more vulnerable to the
harder aluminium-alloy projectile than the soft-gelatine projectile. This is attributed to the
higher contact pressures, and their localisation, associated with the aluminium-alloy projectile.
In the soft impact case, due to the characteristics of the hydrodynamic loading, the impact
energy is mainly absorbed by global bending of the laminates which also gives a longer
contacting time allowing energy spread extensively from the impact zone. Conversely, the due
to locally high stresses and indentation effects, the hard impact case, in which the target
response is highly localised, presents delamination and matrix cracking at initial stage,
followed by the fibre failure, which subsequently leaded to the catastrophic failure in the
composites.

8.5 Modelling unidirectional CFRPs under low-velocity


and high-velocity impact
An elastic, two-dimensional FE model has been developed to simulate the experimental
results obtained from both the low-velocity and high-velocity impact tests. The model, which

130
CHAPTER 8

is relatively computationally efficient, has been shown to simulate the loading responses of the
composites by the impact event and the interlaminar and intralaminar damage induced.

For the low-velocity drop-weight tests, the results from the FE model for the load versus
time curves from all the tests were in very good agreement with the experimental measurements.
Considering the FE modelling of the high-velocity tests, the simulations have accurately
predicted the shape and values of the OOP displacements as a function of the time-scale of the
impact event, that the damage areas in the composites were always significantly higher from
the high-velocity gas-gun tests than from the low-velocity drop-weight tests, at a comparable
impact energy of 10.5 J, and that the CF/Epoxy composite suffered somewhat greater damage
compared to the CF/PEEK composite.

The results from this model has been quantitatively validated against the impact
response of the composites and captures the essential aspects of their impact behaviour and can
be used (a) to simulate aspects of the high-velocity tests that cannot be directly measured and
(b) to optimise the impact performance of such materials in industrial applications.

8.6 Modelling woven-fabric CFRPs under high-velocity


impact
A FE numerical model was developed for predicting the behaviour of the soft gelatine
projectile and the woven-fabric composite test specimen during the impact event. This model
has enabled (a) the deformation, (b) the initiation of damage, and (c) the evolution of such
damage in the composite target specimens to be predicted, as well as the deformation and flow
behaviour (and the contact pressure) of the soft gelatine projectile. The results from the
numerical studies have been found to be in very good agreement with the experimental results.

In terms of design, woven architectures are often employed on the outside of composite
laminates to generate a hybrid architecture. Woven composites do not have the stiffness of an
equivalent laminate carbon-fibre material but they have the advantage that delamination and
interfacial cracking does not occur so readily as has been shown by the modelling and
experiments.

131
Chapter 9. Future work

9.1 Introduction
Carbon fibre reinforced polymers (CFRPs) are, as the most promising materials for
aerospace applications, facing great challenges when encounter with an impact with foreign
bodies. Failure of CFRP structure is often a complex process and require a thorough
understanding. Based on the understanding of the current studies, some further developments
in terms of material characterisation methods, repaired CFRPs and composite damage model
are suggested for further improving the understanding of such materials, and optimaise the
reliability and capability of the developed composite damage model.

9.2 Effects of impact loading on compression-after-


impact (CAI) strength of CFRPs
An impact on a composite structure may result in invisible external damage but a
dramatic degradation of compressive strength. Hence, a compression-after-impact (CAI) test
is a crucial method to evaluate the residual strength of composite materials after impact. Its
values are determined by the resistance to buckling of the sub-laminates, mainly affected by
delamintion, and rated as one of the major screening parameters for material election
[64,179,180].

To extend the existing studies, the thermoplastic and thermoset composites under high-
velocity impact and CAI loading conditions should be investigated. Such residual strength is
an essential factor to be evaluated for critical composite aerostructures. In addition to that, the
thermoplastic and thermoset composite specimens were impacted at a relatively low-velocity
and a high-velocity by using a rigid metallic impactor or a soft gelatine impactor. The effects

132
CHAPTER 9

of such impact, low-velocity impact versus high-velocity impact and soft impact versus hard
impact, on CAI strength are worthy of further investigation.

9.3 Effects of environmental factors on impact


performance of CFRPs
CFRPs are widely implementing on primary structural applications of the latest
commercial aircrafts which operate over a broad range of temperatures from -56.4 °C to 55 °C
at ground level and as low as -80 °C at an altitude of 12,000 m [181–183]. Under such extremes
of the ambient air temperature and thermo-mechanical conditions, the performance of aircrafts
should be established and satisfy the airworthiness certification requirements.

With this in mind, many researchers have studied the impact performance of CFRPs
under impact loading at low and high temperatures [171,184–187]. In a certain extent, low
temperature has led to greater damage extension due to worse dissipation of energy, whereas
the ductile and viscoelastic behaviours of CFRPs, being enhanced at elevated temperature,
weaken their stiffness and residual strength and further influence failure response. In this Ph.D.
project, all tests were performed in a lab environment at constant room temperature and
humidity. Therefore, a better understanding of the impact behaviour of such composite
materials under extreme environmental conditions is needed.

9.4 Performance of the repaired CFRPs under impact


loadings
An area of interest when considering the use of composite materials for aircraft is the
impact performance of repaired composites. The mechanical properties of these materials after
repair are of great interest because the life of an aircraft is significantly longer than some of its
components, meaning repair is critical. It is especially important for composite materials
because the residual strength is significantly reduced after damage, but this drawback can be
mitigated through the use of repair techniques. Currently, there are two main techniques, i.e.
patch repair and scarf joint [188–190], that are well developed. Hence, the impact performace
of the composites repaired via such techniques is of particular interest.

133
CHAPTER 9

As there are a number of types of impact that an aircraft will be subjected to, the
performance of repaired composites under drop-weight impact and gas gun impact are of great
worth to be investigated. In addition, once a damaged sample has been repaired and the
component replaced on the aircraft, it is not often hit in the same place again and so
consideration has to be taken for how the repair performs when the impact is at a distance from
the original impact. Therefore, the impact behaviour of joint area where has been approved as
the weakest part of the repair should be well studied [191,192].

134
List of publication

Journal papers

(1) Liu, H., Liu, J., Ding, Y., Hall, Z.E., Kong, X., Zhou, J., Harper, L., Blackman, B.R.K.,
Kinloch, A.J., Dear, J.P. A three-dimensional elastic-plastic damage model for predicting
the impact behaviour of fibre-reinforced polymer-matrix composites. Compos. B Eng.
2020 (online).

(2) Liu, H., Liu, J., Ding, Y., Zheng, J., Kong, X., Ma, X., Harper, L., Blackman, B.R.K.,
Kinloch, A.J., Dear, J.P. The behaviour of thermoplastic and thermoset carbon-fibre
composites subjected to low-velocity and high-velocity impacts. J. Mater. Sci. 2020;
55:15741-68.

(3) Liu, H., Liu, J., Ding, Y., Zhou, J., Kong, X., Blackman, B.R.K., Kinloch, A.J., Falzon,
B.G., Dear, J.P. Effects of Impactor Geometry on the Low-Velocity Impact Behaviour of
Fibre-Reinforced Composites: An Experimental and Theoretical Investigation. Appl
Compos Mater. 2020 (online).

(4) Liu, H., Liu, J., Kaboglu, C., Zhou, J., Kong, X., Blackman, B.R.K., Kinloch, A.J., Dear,
J.P. The behaviour of fibre-reinforced composites subjected to a soft impact-loading: An
experimental and numerical study. Eng Fail Anal. 2020;111:104448.

(5) Liu, H., Liu, J., Ding, Y., Zhou, J., Kong, X., Harper, L.T., Blackman, B.R.K., Falzon,
B.G., Dear, J.P. Modelling damage in fibre-reinforced thermoplastic composite laminates
subjected to three-point bend loading. Compos Struct. 2020;236:111889.

(6) Liu, H., Liu, J., Kaboglu, C., Chai, H., Kong, X., Blackman, B.R.K., Kinloch, A.J., Dear,
J.P. Experimental investigations on the effects of projectile hardness on the impact
response of fibre-reinforced composite laminates. Int J Light Mater Manuf. 2020;3:77–
87.

135
LIST OF PUBLICATION

(7) Liu, J., Liu, H., Kaboglu, C., Kong, X., Ding, Y., Chai, H., Blackman, B.R.K., Kinloch,
A.J., Dear, J.P. The Impact Performance of Woven-Fabric Thermoplastic and Thermoset
Composites Subjected to High-Velocity Soft- and Hard-Impact Loading. Appl Compos
Mater. 2019;26:1389–410.

(8) Liu, H., Liu, J., Kaboglu, C., Chai, H., Kong, X., Blackman, B.R.K., Kinloch, A.J., Dear,
J.P. Experimental and numerical studies on the behaviour of fibre-reinforced composites
subjected to soft impact loading. Procedia Struct Integr. 2019;17:992–1001.

(9) Domun, N., Kaboglu, C., Paton, K.R., Dear, J.P., Liu, J., Blackman, B.R.K., Liaghat, G.,
Hadavinia, H. Ballistic impact behaviour of glass fibre reinforced polymer composite with
1D/2D nanomodified epoxy matrices. Compos Part B Eng. 2019;167:497–506.

(10) Zhou, J., Liu, J., Zhang, X., Yan, Y., Jiang, L., Mohagheghian, I., Dear, J.P.,
Charalambides, M.N. Experimental and numerical investigation of high velocity soft
impact loading on aircraft materials. Aerosp Sci Technol. 2019;90:44–58.

(11) Liu, H., Falzon, B.G., Li, S., Tan, W., Liu, J., Chai, H., Blackman, B.R.K., Dear, J.P.
Compressive failure of woven fabric reinforced thermoplastic composites with an open-
hole: An experimental and numerical study. Compos Struct. 2019;213:108–17.

Conference papers

(12) Liu, J., Liu, H., Chai, H., Harper, L.T., Falzon, B.G., Blackman, B.R.K., Kinloch, A.J.,
Dear, J.P. The performance of thermoplastic and thermoset composites subjected to low-
velocity and high-velocity impact loading. In: 22nd International Conference on
Composite Materials. Melbourne, Australia; 2019.

(13) Liu, J., Kaboglu, C., Liu, H., Blackman, B.R.K., Kinloch, A.J., Dear, J.P. High speed
digital image correlation for impact performance of thermoplastic and thermoset
composites. In: 18th European Conference on Composite Materials. Athen, Greece; 2018.

136
Bibliography

[1] Aerospace Technology Institute. Composite material applications in aerospace.


Aerosp Technol Inst. 2018;1–12.

[2] Newaz, G.M. Advances in Thermoplastic Matrix Composite Materials. ASTM


International, USA; 1989.

[3] Mallick, P. Fibre-reinforced composites materials, manufacturing and design. 3rd ed.
CRC Press; 2007.

[4] Strong, A.B. Fundamentals of composites manufacturing: materials, methods and


applications. Society of Manufacturing Engineers; 2008.

[5] Cogswell, F.N. Thermoplastic aromatic polymer composites: a study of the structure,
processing and properties of carbon fibre reinforced polyetheretherketone and related
materials. Elsevier, UK; 2013.

[6] Long, A.C. Composites forming technologies. Elsevier, UK; 2014.

[7] Ngo, T.-D. Introduction to Composite Materials. In: Fiber Composites. 2020. p. 27.

[8] Wennberg, D. Multi-Functional Composite Design Concepts for Rail Vehicle Car
Bodies. PhD Thesis; 2013.

[9] McLAREN Group. McLAREN MCL34 Technical Specification. 2019.

[10] Hellard, G. Composites in Airbus: A long story of innovations and experiences. In:
Airbus Global Investor Forum. Seville, Spain; 2008.

[11] Hawk, J. The Boeing 787 Dreamliner: More than an airplane. In: Aircraft Noise and
Emissions Reduction Symposium. Monterey, USA; 2005.

137
BIBLIOGRAPHY

[12] Foster, D.S., Applman, H.S. Aviation hail problem (WMO-No. 109. TP. 47). Geneva,
Switzerland: World Meteorological Organization; 1961.

[13] Herricks, E., Mayer, D., Majumdar, S. Foreign object debris characterization at a
large international airport. Springfield, USA; 2015.

[14] Alves, M., Chaves, C.E., Birch, R.S. Impact on aircraft. 17th Int Congr Mech Eng.
2003;1–8.

[15] Dolbeer, R.A., Begier, M.J., Miller, P.R., Weller, J.R., Anderson, A.L. Wildlife strikes
to civil aircraft in the United States 1990-2018. Federal Aviation Administration,
USA; 2019.

[16] Bennett, D.L. Debris hazards at airports. Federal Aviation Adiministration, USA;
1996.

[17] Samenow, J. ‘Scariest flight of my life’: Hail smashes nose of plane that flew into
towering storm. The Washington Post. 2018;

[18] John Hutchinson. Incredible photos reveal how aircraft’s nose collapsed after bird
flew into Turkish Airlines plane carrying 125 passengers. Daily Mail. 2015;

[19] Bachtel, B. Foreign Object Debris and Damage Prevention. The Boeing Company,
USA; 2016.

[20] Kaw, A.K. Mechanics of composite materials. CRC press; 2005.

[21] Matthews, F.L. Composite materials : engineering and science. Rawlings RD, editor.
Woodhead Publishing Series in Composites Science and Engineering. Burlington:
Elsevier Science; 1999.

[22] Games, S. Method and apparatus for making glass wool. 1938.

[23] Heine, M. Carbon Fibers. In: Jager H, Frohs W, editors. Industrial Carbon and
Graphite Materials. 2020.

[24] Chatzi, E.G., Koenig, J.L. Morphology and Structure of Kevlar Fibers: A Review.
Polym Plast Technol Eng. 1987;26(3–4):229–70.

[25] Tong, L., Mouritz, A.P., Bannister, M.K. 3D Fibre Reinfored Polymer Composites.
Oxford: Elsevier Science; 2002.

138
BIBLIOGRAPHY

[26] Nelson, J.W. Composite Materials for Aircraft Structures: A Brief Review of Practical
Application. Bozeman,Montana,USA: Montana State University; 2010.

[27] Campbell, F.C. Structural composite materials. ASM international; 2010.

[28] Bergstrom, J.S. Mechanics of solid polymers: theory and computational modeling.
William Andrew; 2015.

[29] Ebewele, R.O. Polymer science and technology. Polymer Science and Technology.
2000. 1–531 p.

[30] Blundell, D.J., Chalmers, J.M., Mackenzie, M.W., Gaskin, W.F. Crystalline
morphology of the matrix of PEEK-carbon fiber aromatic polymer composites. I.
Assessment of crystallinity.[Polyetheretherketone]. SAMPE Q;(United States).
1985;16(4).

[31] Blundell, D.J., Crick, R.A., Fife, B., Peacock, J., Keller, A., Waddon, A. Spherulitic
morphology of the matrix of thermoplastic PEEK/carbon fibre aromatic polymer
composites. J Mater Sci. 1989;24(6):2057–64.

[32] Saliba, T.E., Anderson, D.P., Servais, R.A. Process modeling of heat transfer and
crystallization in complex shapes thermoplastic composites. J Thermoplast Compos
Mater. 1989;2(2):91–104.

[33] Leicy, D., Hogg, P.J. The effect of crystallinity on the impact properties of advanced
thermoplastic composites. In: Developments in the Science and Technology of
Composite Materials. Springer; 1989. p. 809–15.

[34] Fujihara, K., Huang, Z.-M., Ramakrishna, S., Hamada, H. Influence of processing
conditions on bending property of continuous carbon fiber reinforced PEEK
composites. Compos Sci Technol. 2004;64(16):2525–34.

[35] Jar, P.-Y., Mulone, R., Davies, P., Kausch, H.-H. A study of the effect of forming
temperature on the mechanical behaviour of carbon-fibre/peek composites. Compos
Sci Technol. 1993;46(1):7–19.

[36] Fernlund, G., Rahman, N., Courdji, R., Bresslauer, M., Poursartip, A., Willden, K.,
Nelson, K. Experimental and numerical study of the effect of cure cycle, tool surface,
geometry, and lay-up on the dimensional fidelity of autoclave-processed composite
parts. Compos part A Appl Sci Manuf. 2002;33(3):341–51.

139
BIBLIOGRAPHY

[37] Kim, Y.K., Daniel, I.M. Cure cycle effect on composite structures manufactured by
resin transfer molding. J Compos Mater. 2002;36(14):1725–43.

[38] Coenen, V., Hatrick, M., Law, H., Brosius, D., Nesbitt, A., Bond, D. A feasibility study
of Quickstep Processing of an aerospace composite material. In: SAMPE Europe.
Paris, France; 2005.

[39] Davies, L.W., Day, R.J., Bond, D., Nesbitt, A., Ellis, J., Gardon, E. Effect of cure cycle
heat transfer rates on the physical and mechanical properties of an epoxy matrix
composite. Compos Sci Technol. 2007;67(9):1892–9.

[40] Olabi, A.G., Lostado-Lorza, R., Benyounis, K. Review of Microstructures, Mechanical


Properties, and Residual Stresses of Ferritic and Martensitic Stainless-Steel Welded
Joints. In: Comprehensive Materials Processing. 2014. p. 181–92.

[41] Cantwell, W.J., Morton, J. The impact resistance of composite materials — a review.
Composites. 1991;22(5):347–62.

[42] Richardson, M.O.W., Wisheart, M.J. Review of low-velocity impact properties of


composite materials. Compos Part A Appl Sci Manuf. 1996;27(12):1123–31.

[43] Davies, G.A.O., Olsson, R. Impact on composite structures. Aeronaut J.


2004;108(1089):541–63.

[44] Jogur, G., Nawaz Khan, A., Das, A., Mahajan, P., Alagirusamy, R. Impact properties
of thermoplastic composites. Text Prog. 2018;50(3):109–83.

[45] Robinson, P., Davies, G.A.O. Impactor mass and specimen geometry effects in low
velocity impact of laminated composites. Int J Impact Eng. 1992;12(2):189–207.

[46] Cantwell, W.J., Morton, J. Comparison of the low and high velocity impact response of
CFRP. Composites. 1989;20(6):545–51.

[47] Standard Test Method for Measuring the Damage Resistance of a Fiber-Reinforced
Polymer Matrix Composite to a Drop-Weight Impact Event (D7136/D7136M−15).
West Conshohocken, PA, USA; 2015.

[48] Cantwell, W.J., Morton, J. Detection of impact damage in CFRP laminates. Compos
Struct. 1985;3(3–4):241–57.

140
BIBLIOGRAPHY

[49] Dorey, G., Bishop, S.M., Curtis, P.T. On the impact performance of carbon fibre
laminates with epoxy and PEEK matrices. Compos Sci Technol. 1985;23(3):221–37.

[50] Bishop, S.M. The mechanical performance and impact behaviour of carbon-fibre
reinforced PEEK. Compos Struct. 1985;3(3–4):295–318.

[51] Vieille, B., Casado, V.M., Bouvet, C. About the impact behavior of woven-ply carbon
fiber-reinforced thermoplastic- and thermosetting-composites : A comparative study.
Compos Struct. 2013;101:9–21.

[52] Chamis, C.C. Designing for impact resistance with unidirectional fiber composites.
National Aeronautics and Space Administration; 1971.

[53] Sundaram, A.S., Eranezhuth, A.A., Krishna, K., Kumar, P.K., Sivakumar, V. Ballistic
impact performance study on thermoset and thermoplastic composites. J Fail Anal
Prev. 2017;17(6):1260–7.

[54] Faur‐Csukat, G. A study on the ballistic performance of composites. In:


Macromolecular symposia. Wiley Online Library; 2006. p. 217–26.

[55] Will, M.A., Franz, T., Nurick, G.N. The effect of laminate stacking sequence of CFRP
filament wound tubes subjected to projectile impact. Compos Struct. 2002;58(2):259–
70.

[56] Abrate, S. Impact on Composite Structures. Cambridge: Cambridge University Press;


1998.

[57] Raimondo, L., Iannucci, L., Robinson, P., Pinho, S.T., Curtis, P.T., Wells, G.M.
Predicting the dynamic behaviour of polymer composites. ICCM Int Conf Compos
Mater. 2007;1–10.

[58] Higuchi, R., Okabe, T., Yoshimura, A., Tay, T.E. Progressive failure under high-
velocity impact on composite laminates: Experiment and phenomenological
mesomodeling. Eng Fract Mech. 2017;178:346–61.

[59] Morita, H., Adachi, T., Tateishi, Y., Matsumot, H. Characterization of impact damage
resistance of cf/peek and cf/toughened epoxy laminates under low and high velocity
impact tests. J Reinf Plast Compos. 1997;16(2):131–43.

141
BIBLIOGRAPHY

[60] Wagner, T., Heimbs, S., Franke, F., Burger, U., Middendorf, P. Experimental and
numerical assessment of aerospace grade composites based on high-velocity impact
experiments. Compos Struct. 2018;204(July):142–52.

[61] Husman, G.E., Whitney, J.M., Halpin, J.C. Residual strength characterization of
laminated composites subjected to impact loading. In: Foreign object impact damage
to composites. ASTM International; 1975.

[62] Nash, N.H., Young, T.M., McGrail, P.T., Stanley, W.F. Inclusion of a thermoplastic
phase to improve impact and post-impact performances of carbon fibre reinforced
thermosetting composites—A review. Mater Des. 2015;85:582–97.

[63] Murugan, P., Naresh, K., Shankar, K., Velmurugan, R., Balaganesan, G. High velocity
impact damage investigation of carbon/epoxy/clay nanocomposites using 3D
Computed Tomography. Mater Today Proc. 2018;5(9):16946–55.

[64] Vieille, B., Casado, V.M., Bouvet, C. Influence of matrix toughness and ductility on
the compression-after-impact behavior of woven-ply thermoplastic- and thermosetting-
composites: A comparative study. Compos Struct. 2014;110:207–18.

[65] Bull, D.J., Scott, A.E., Spearing, S.M., Sinclair, I. The influence of toughening-
particles in CFRPs on low velocity impact damage resistance performance. Compos
Part A Appl Sci Manuf. 2014;58:47–55.

[66] Cantwell, W.J., Morton, J. Impact perforation of carbon fibre reinforced plastic.
Compos Sci Technol. 1990;38(2):119–41.

[67] Wang, B., Xiong, J., Wang, X., Ma, L., Zhang, G.Q., Wu, L.Z., Feng, J.C. Energy
absorption efficiency of carbon fiber reinforced polymer laminates under high velocity
impact. Mater Des. 2013;50:140–8.

[68] Appleby-Thomas, G.J., Wood, D.C., Hameed, A., Leighs, J.A. On the ballistic
response of an aerospace-grade composite panel to non-spheroidised fragment
simulants. Compos Struct. 2015;119:90–8.

[69] Babu, M.G., Velmurugan, R., Gupta, N.K. Energy-absorption capability of thin
laminates subjected to heavy-mass projectile impact of varying nose geometries. Int J
Crashworthiness. 2008;13(3):237–46.

142
BIBLIOGRAPHY

[70] Seifoori, S., Izadi, R., Yazdinezhad, A.R. Impact damage detection for small- and
large-mass impact on CFRP and GFRP composite laminate with different striker
geometry using experimental, analytical and FE methods. Acta Mech.
2019;230(12):4417–33.

[71] Daniel, I.M., Werner, B.T., Fenner, J.S. Strain-rate-dependent failure criteria for
composites. Compos Sci Technol. 2011;71(3):357–64.

[72] Kline, R.A., Chang, F.H. Composite Failure Surface Analysis. J Compos Mater.
1980;14(4):315–24.

[73] Hinton, M.J., Soden, P.D. Predicting failure in composite laminates: the background
to the exercise. Compos Sci Technol. 1998;58(7):1001–10.

[74] Laffan, M.J., Pinho, S.T., Robinson, P., McMillan, A.J. Translaminar fracture
toughness testing of composites: A review. Polym Test. 2012;31(3):481–9.

[75] Pinho, S.T., Iannucci, L., Robinson, P. Physically based failure models and criteria for
laminated fibre-reinforced composites with emphasis on fibre kinking. Part II: FE
implementation. Compos Part A Appl Sci Manuf. 2006;37(5):766–77.

[76] Williams, K. V, Vaziri, R., Poursartip, A. A physically based continuum damage


mechanics model for thin laminated composite structures. Int J Solids Struct.
2003;40(9):2267–300.

[77] Iannucci, L., Willows, M.L. An energy based damage mechanics approach to
modelling impact onto woven composite materials: Part II. Experimental and
numerical results. Compos Part A Appl Sci Manuf. 2007;38(2):540–54.

[78] He, Y., Makeev, A., Shonkwiler, B. Characterization of nonlinear shear properties for
composite materials using digital image correlation and finite element analysis.
Compos Sci Technol. 2012;73:64–71.

[79] Giannadakis, K., Varna, J. Analysis of nonlinear shear stress–strain response of


unidirectional GF/EP composite. Compos Part A Appl Sci Manuf. 2014;62:67–76.

[80] Joki, R.K., Grytten, F., Hayman, B. Nonlinear response in glass fibre non-crimp fabric
reinforced vinylester composites. Compos Part B Eng. 2015;77:105–11.

[81] Pagano, N.J. Interlaminar response of composite materials. Elsevier; 2012.

143
BIBLIOGRAPHY

[82] Wisnom, M.R. The role of delamination in failure of fibre-reinforced composites.


Philos Trans R Soc A Math Phys Eng Sci. 2012;370(1965):1850–70.

[83] Hashemi, S., Kinloch, A.J., Williams, J.M., Ford, H. The analysis of interlaminar
fracture in uniaxial fibre-polymer composites. Proc R Soc London A Math Phys Sci.
1990;427(1872):173–99.

[84] Camanho, P., Dávila, C., De Moura, M. Numerical Simulation of Mixed-Mode


Progressive Delamination in Composite Materials. J Compos Mater. 2003;37.

[85] Goyal, V.K., Jaunky, N.R., Johnson, E.R., Ambur, D.R. Intralaminar and interlaminar
progressive failure analyses of composite panels with circular cutouts. Compos Struct.
2004;64(1):91–105.

[86] Davila, C.G., Camanho, P.P., Rose, C.A. Failure criteria for FRP laminates. J
Compos Mater. 2005;39(4):323–45.

[87] Chowdhury, N.T., Wang, J., Chiu, W.K., Yan, W. Matrix failure in composite
laminates under tensile loading. Compos Struct. 2016;135:61–73.

[88] Herrera-Franco, P.J., Valadez-González, A. Mechanical properties of continuous


natural fibre-reinforced polymer composites. Compos Part A Appl Sci Manuf.
2004;35(3):339–45.

[89] Green, B.G., Wisnom, M.R., Hallett, S.R. An experimental investigation into the
tensile strength scaling of notched composites. Compos Part A Appl Sci Manuf.
2007;38(3):867–78.

[90] Laval, C. Composites design in the real world. Reinforced Plastics. 2003;47(8):50–3.

[91] Daniel, I.M., Luo, J.J., Schubel, P.M. Three-dimensional characterization of textile
composites. Compos Part B Eng. 2008;39(1):13–9.

[92] Tehrani, M., Boroujeni, A.Y., Hartman, T.B., Haugh, T.P., Case, S.W., Al-Haik, M.S.
Mechanical characterization and impact damage assessment of a woven carbon fiber
reinforced carbon nanotube-epoxy composite. Compos Sci Technol. 2013;75:42–8.

[93] Ng, W.H., Salvi, A.G., Waas, A.M. Characterization of the in-situ non-linear shear
response of laminated fiber-reinforced composites. Compos Sci Technol.
2010;70(7):1126–34.

144
BIBLIOGRAPHY

[94] París, F., Blázquez, A., McCartney, L.N., Barroso, A. Characterization and evolution
of matrix and interface related damage in [0/90]S laminates under tension. Part II:
Experimental evidence. Compos Sci Technol. 2010;70(7):1176–83.

[95] Camanho, P.P., Dávila, C.G., Pinho, S.T., Iannucci, L., Robinson, P. Prediction of in
situ strengths and matrix cracking in composites under transverse tension and in-plane
shear. Compos Part A Appl Sci Manuf. 2006;37(2):165–76.

[96] Bogetti, T.A., Staniszewski, J., Burns, B.P., Hoppel, C.P.R., Gillespie, J.W., Tierney,
J. Predicting the nonlinear response and progressive failure of composite laminates
under tri-axial loading. J Compos Mater. 2012;46(19–20):2443–59.

[97] Van Paepegem, W., De Baere, I., Degrieck, J. Modelling the nonlinear shear stress-
strain response of glass fibre-reinforced composites. Part II: Model development and
finite element simulations. Compos Sci Technol. 2006;66(10):1465–78.

[98] Fanteria, D., Panettieri, E. A non-linear model for in-plane shear damage and failure
of composite laminates. Aerotec Missili Spaz. 2014;93(1–2):17–24.

[99] Tan, W., Falzon, B.G. Modelling the nonlinear behaviour and fracture process of
AS4/PEKK thermoplastic composite under shear loading. Compos Sci Technol.
2016;126:60–77.

[100] Fawaz, Z. Quality control and testing methods for advanced composite materials in
aerospace engineering. In: Advanced Composite Materials for Aerospace Engineering.
Elsevier; 2016. p. 429–51.

[101] Okafor, A.C., Otieno, A.W., Dutta, A., Rao, V.S. Detection and characterization of
high-velocity impact damage in advanced composite plates using multi-sensing
techniques. Compos Struct. 2001;54(2–3):289–97.

[102] Cantwell, W.J., Morton, J. The influence of varying projectile mass on the impact
response of CFRP. Compos Struct. 1989;13(2):101–14.

[103] Garcea, S.C., Wang, Y., Withers, P.J. X-ray computed tomography of polymer
composites. Compos Sci Technol. 2018;156:305–19.

[104] Pelivanov, I., Ambroziński, Ł., Khomenko, A., Koricho, E.G., Cloud, G.L., Haq, M.,
O’Donnell, M. High resolution imaging of impacted CFRP composites with a fiber-
optic laser-ultrasound scanner. Photoacoustics. 2016;4(2):55–64.

145
BIBLIOGRAPHY

[105] Li, Z., Meng, Z. A review of the radio frequency non-destructive testing for carbon-
fibre composites. Meas Sci Rev. 2016;16(2):68–76.

[106] Mook, G., Lange, R., Koeser, O. Non-destructive characterisation of carbon-fibre-


reinforced plastics by means of eddy-currents. Compos Sci Technol. 2001;61(6):865–
73.

[107] Gholizadeh, S. A review of non-destructive testing methods of composite materials.


Procedia Struct Integr. 2016;1:50–7.

[108] Oguibe, C.N., Webb, D.C. Finite-element modelling of the impact response of a
laminated composite plate. Compos Sci Technol. 1999;59(12):1913–22.

[109] Shi, Y., Swait, T., Soutis, C. Modelling damage evolution in composite laminates
subjected to low velocity impact. Compos Struct. 2012;94(9):2902–13.

[110] Johnson, A.F., Holzapfel, M. Modelling soft body impact on composite structures.
Compos Struct. 2003;61(1–2):103–13.

[111] Liu, H., Tang, Z., Pan, L., Zhao, W., Sun, B., Jiang, W. Numerical simulating and
experimental study on the woven carbon fiber-reinforced composite laminates under
low-velocity impact. In: AIP Conference Proceedings. AIP Publishing LLC; 2016. p.
20110.

[112] Evci, C., Gülgeç, M. An experimental investigation on the impact response of


composite materials. Int J Impact Eng. 2012;43:40–51.

[113] Kurşun, A., Şenel, M., Enginsoy, H.M., Bayraktar, E. Effect of impactor shapes on the
low velocity impact damage of sandwich composite plate: Experimental study and
modelling. Compos Part B Eng. 2016;86:143–51.

[114] Tita, V., de Carvalho, J., Vandepitte, D. Failure analysis of low velocity impact on thin
composite laminates: Experimental and numerical approaches. Compos Struct.
2008;83(4):413–28.

[115] Rybicki, E.F., Kanninen, M.F. A finite element calculation of stress intensity factors by
a modified crack closure integral. Eng Fract Mech. 1977;9(4):931–8.

[116] de Moura, M.F.S.F., Silva, M.A.L., de Morais, A.B., Morais, J.J.L. Equivalent crack
based mode II fracture characterization of wood. Eng Fract Mech. 2006;73(8):978–93.

146
BIBLIOGRAPHY

[117] de Morais, A.B., Pereira, A.B. Mixed mode I+II interlaminar fracture of glass/epoxy
multidirectional laminates – Part 1: Analysis. Compos Sci Technol.
2006;66(13):1889–95.

[118] Barenblatt, G.I. The Mathematical Theory of Equilibrium Cracks in Brittle Fracture.
Adv Appl Mech. 1962;7(C):55–129.

[119] Camanho, P.P., Dávila, C.G. Mixed-mode decohesion finite elements for the simulation
of delamination in composite materials. 2002;

[120] Whitney, J.M., Browning, C.E. Materials Characterization for Matrix-Dominated


Failure Modes. ASTM Spec Tech Publ. 1984;104–24.

[121] Borg, R., Nilsson, L., Simonsson, K. Simulation of low velocity impact on fiber
laminates using a cohesive zone based delamination model. Compos Sci Technol.
2004;64(2):279–88.

[122] González, E. V, Maimí, P., Martín-Santos, E., Soto, A., Cruz, P., Martín de la
Escalera, F., Sainz de Aja, J.R. Simulating drop-weight impact and compression after
impact tests on composite laminates using conventional shell finite elements. Int J
Solids Struct. 2018;144–145:230–47.

[123] Liao, B.B., Liu, P.F. Finite element analysis of dynamic progressive failure of plastic
composite laminates under low velocity impact. Compos Struct. 2017;159:567–78.

[124] Hongkarnjanakul, N., Bouvet, C., Rivallant, S. Validation of low velocity impact
modelling on different stacking sequences of CFRP laminates and influence of fibre
failure. Compos Struct. 2013;106:549–59.

[125] Puck, A., Schürmann, H. Failure analysis of FRP laminates by means of physically
based phenomenological models. Compos Sci Technol. 2002;62(12–13):1633–62.

[126] Dvorak, G.J., Laws, N. Analysis of progressive matrix cracking in composite laminates
II. First ply failure. J Compos Mater. 1987;21(4):309–29.

[127] Liu, H., Falzon, B.G., Tan, W. Predicting the Compression-After-Impact (CAI)
strength of damage-tolerant hybrid unidirectional/woven carbon-fibre reinforced
composite laminates. Compos Part A Appl Sci Manuf. 2018;105:189–202.

147
BIBLIOGRAPHY

[128] Matzenmiller, A., Lubliner, J., Taylor, R.L. A constitutive model for anisotropic
damage in fiber-composites. Mech Mater. 1995;20(2):125–52.

[129] Hallquist, J. LS-DYNA theory manual. Livermore Software Technology Corporation;


2006.

[130] Chang, F.-K., Chang, K.-Y. A progressive damage model for laminated composites
containing stress concentrations. J Compos Mater. 1987;21(9):834–55.

[131] Shi, Y., Pinna, C., Soutis, C. Modelling impact damage in composite laminates: A
simulation of intra- and inter-laminar cracking. Compos Struct. 2014;114(1):10–9.

[132] Ansari, M.M., Chakrabarti, A. Impact behavior of FRP composite plate under low to
hyper velocity impact. Compos Part B Eng. 2016;95:462–74.

[133] Donadon, M. V, Falzon, B.G., Iannucci, L., Hodgkinson, J.M. Measurement of Fibre
Fracture Toughness Using an Alternative Specimen Geometry. Proc 16th Int Conf
Compos Mater July 2007, Kyoto, Japan. 2007;

[134] Donadon, M. V, De Almeida, S.F.M., Arbelo, M.A., de Faria, A.R. A three-


dimensional ply failure model for composite structures. Int J Aerosp Eng. 2009;2009.

[135] Faggiani, A., Falzon, B.G. Predicting low-velocity impact damage on a stiffened
composite panel. Compos Part A Appl Sci Manuf. 2010;41:737–49.

[136] Falzon, B.G., Apruzzese, P. Numerical analysis of intralaminar failure mechanisms in


composite structures. Part I: FE implementation. Compos Struct. 2011;93(2):1039–46.

[137] Falzon, B.G., Apruzzese, P. Numerical analysis of intralaminar failure mechanisms in


composite structures. Part II: Applications. Compos Struct. 2011;93(2):1047–53.

[138] Nishikawa, M., Hemmi, K., Takeda, N. Finite-element simulation for modeling
composite plates subjected to soft-body, high-velocity impact for application to bird-
strike problem of composite fan blades. Compos Struct. 2011;93(5):1416–23.

[139] Liu, J., Li, Y., Gao, X. Bird strike on a flat plate: Experiments and numerical
simulations. Int J Impact Eng. 2014;70:21–37.

[140] Cantwell, W.J., Curtis, P.T., Morton, J. An assessment of the impact performance of
CFRP reinforced with high-strain carbon fibres. Compos Sci Technol.
1986;25(2):133–48.

148
BIBLIOGRAPHY

[141] Cytec Industries Inc. APC-2-PEEK Thermoplastic Polymer Technical Data Sheet.
2015;

[142] Toray Composite Materials America Inc. T700S Standard modulus carbon fiber. 2018;

[143] GOM mbH. Aramis User Manual. 1301_01_EN. Braunschweig, Germany; 2016. 1–57
p.

[144] Williams, J.G., Adams, G.C. The analysis of instrumented impact tests using a mass-
spring model. Int J Fract. 1987;33(3):209–22.

[145] Crouch, B.A., Williams, J.G. Modelling of dynamic crack propagation behaviour in
the three-point bend impact specimen. J Mech Phys Solids. 1988;36(1):1–13.

[146] Dear, J.P., MacGillivray, J.H. Strain gauging for accurate determination of K and G in
impact tests. J Mater Sci. 1991;26(8):2124–32.

[147] Liu, H., Falzon, B.G., Tan, W. Experimental and numerical studies on the impact
response of damage-tolerant hybrid unidirectional/woven carbon-fibre reinforced
composite laminates. Compos Part B Eng. 2018;136(October 2017):101–18.

[148] Abaqus 2018 documentation. Dassault Systèmes. Provid Rhode Island, USA. 2018;

[149] Tan, W., Falzon, B.G. Modelling the crush behaviour of thermoplastic composites.
Compos Sci Technol. 2016;134:57–71.

[150] Chiu, L.N.S., Falzon, B.G., Chen, B., Yan, W. Validation of a 3D damage model for
predicting the response of composite structures under crushing loads. Compos Struct.
2016;147:65–73.

[151] Hashin, Z., Rotem, A. A Fatigue Failure Criterion for Fiber Reinforced Materials. J
Compos Mater. 1973;7:448–64.

[152] Hashin, Z. Failure criteria for unidirectional fiber composites. J Appl Mech.
1980;47(2):329–34.

[153] Liu, H., Falzon, B.G., Li, S., Tan, W., Liu, J., Chai, H., Blackman, B.R.K., Dear, J.P.
Compressive failure of woven fabric reinforced thermoplastic composites with an
open-hole: An experimental and numerical study. Compos Struct. 2019;213:108–17.

[154] Dugdale, D.S. Yielding of steel sheets containing slits. J Mech Phys Solids.
1960;8(2):100–4.
149
BIBLIOGRAPHY

[155] Turon, A., Dávila, C.G., Camanho, P.P., Costa, J. An engineering solution for mesh
size effects in the simulation of delamination using cohesive zone models. Eng Fract
Mech. 2007;74(10):1665–82.

[156] Shi, Y., Pinna, C., Soutis, C. Interface cohesive elements to model matrix crack
evolution in composite laminates. Appl Compos Mater. 2014;21(1):57–70.

[157] Brewer, J.C., Lagace, P.A. Quadratic Stress Criterion for Initiation of Delamination. J
Compos Mater. 1988;22(12):1141–55.

[158] Benzeggagh, M.L., Kenane, M. Measurement of mixed-mode delamination fracture


toughness of unidirectional glass/epoxy composites with mixed-mode bending
apparatus. Compos Sci Technol. 1996;56(4):439–49.

[159] Sarrado, C., Turon, A., Renart, J., Urresti, I. Assessment of energy dissipation during
mixed-mode delamination growth using cohesive zone models. Compos Part A Appl
Sci Manuf. 2012;43(11):2128–36.

[160] Kim, J.-K.K., Sham, M.-L.L. Impact and delamination failure of woven-fabric
composites. Compos Sci Technol. 2000;60(5):745–61.

[161] Iannucci, L., Willows, M.L. An energy based damage mechanics approach to
modelling impact onto woven composite materials—Part I: Numerical models.
Compos Part A Appl Sci Manuf. 2006;37(11):2041–56.

[162] Turon, A., Camanho, P.P., Costa, J., Renart, J. Accurate simulation of delamination
growth under mixed-mode loading using cohesive elements: Definition of interlaminar
strengths and elastic stiffness. Compos Struct. 2010;92(8):1857–64.

[163] Naderi, M., Khonsari, M.M. Stochastic analysis of inter-and intra-laminar damage in
notched PEEK laminates. Exp Polym Lett. 2013;7:383–95.

[164] Abir, M.R., Tay, T.E., Ridha, M., Lee, H.P. On the relationship between failure
mechanism and compression after impact (CAI) strength in composites. Compos
Struct. 2017;182(April 2018):242–50.

[165] Blackman, B.R.K., Dear, J.P., Kinloch, A.J., Macgillivray, H., Wang, Y., Williams,
J.G., Yayla, P. The failure of fibre composites and adhesively bonded fibre composites
under high rates of test - Part I Mode I loading-experimental studies. J Mater Sci.
1995;30(23):5885–900.

150
BIBLIOGRAPHY

[166] Blackman, B.R.K., Kinloch, A.J., Wang, Y., Williams, J.G. The failure of fibre
composites and adhesively bonded fibre composites under high rates of test: Part II
mode I loading - Dynamic effects. J Mater Sci. 1996;31(17):4451–66.

[167] Blackman, B.R.K., Dear, J.P., Kinloch, A.J., MacGillivray, H., Wang, Y., Williams,
J.G., Yayla, P. The failure of fibre composites and adhesively bonded fibre composites
under high rates of test: Part III mixed-mode I/II and mode II loadings. J Mater Sci.
1996;31(17):4467–77.

[168] Kinloch, A.J., Kodokian, G.A., Jamarani, M.B. Impact properties of epoxy polymers. J
Mater Sci. 1987;22(11):4111–20.

[169] Dear, J.P. High-speed photography of impact effects in three-point bend testing of
polymers. J Appl Phys. 1990;67(9):4304–12.

[170] Wilbeck, J.S., Rand, J.L. The development of a substitute bird model. J Eng Power.
1981;103(4):725–30.

[171] López-Puente, J., Zaera, R., Navarro, C. The effect of low temperatures on the
intermediate and high velocity impact response of CFRPs. Compos Part B Eng.
2002;33(8):559–66.

[172] Hashin, Z. Failure Criteria for Unidirectional Fiber Composites. J Appl Mech.
2015;47:329–34.

[173] Liu, J., Liu, H., Kaboglu, C., Kong, X., Ding, Y., Chai, H., Blackman, B.R.K.,
Kinloch, A.J., Dear, J.P. The Impact Performance of Woven-Fabric Thermoplastic and
Thermoset Composites Subjected to High-Velocity Soft- and Hard-Impact Loading.
Appl Compos Mater. 2019;26:1389–410.

[174] Liu, H., Falzon, B.G., Catalanotti, G., Tan, W. An experimental method to determine
the intralaminar fracture toughness of high-strength carbon-fibre reinforced
composite aerostructures. Aeronaut J. 2018;122(1255):1352–70.

[175] Falzon, B.G., Liu, H., Tan, W. Comment on A tensorial based progressive damage
model for fiber reinforced polymers. Compos Struct. 2017;176:877–82.

[176] Radchenko, A., Radchenko, P. Numerical modeling of development of fracture in


anisotropic composite materials at low-velocity loading. J Mater Sci.
2011;46(8):2720–5.

151
BIBLIOGRAPHY

[177] CYTEC. APC-2-PEEK Thermoplastic Polymer. 2012.

[178] Naderi, M., Khonsari, M.M. Stochastic analysis of inter- and intra-laminar damage in
notched PEEK laminates. Express Polym Lett. 2013;7(4):383–95.

[179] Ishikawa, T., Sugimoto, S., Matsushima, M., Hayashi, Y. Some experimental findings
in compression-after-impact (CAI) tests of CF/PEEK (APC-2) and conventional
CF/epoxy flat plates. Compos Sci Technol. 1995;55(4):349–63.

[180] Tuo, H., Lu, Z., Ma, X., Xing, J., Zhang, C. Damage and failure mechanism of thin
composite laminates under low-velocity impact and compression-after-impact loading
conditions. Compos Part B Eng. 2019;163:642–54.

[181] European Union Aviation Safety Agency. Easy Access Rules for Airworthiness and
Environmental Certification (Regulation (EU) No 748/2012). 2019.

[182] Federal Aviation Administration. Flight Test Guide for Certification of Transport
Category Airplanes. Area. 2018.

[183] Lanson, F., von Wrede, R. The Aircraft Environmental Flight Envelope. In: WMO
Aeronautical Meteorology Scientific Conference (AMSC-2017). 2017. p. 1–11.

[184] Bibo, G. a., Hogg, P.J., Kemp, M. High-temperature damage tolerance of carbon
fibre-reinforced plastics: Composites. 1995;26(2):91–102.

[185] Tai, N.-H., Yip, M.-C., Tseng, C.-M. Influences of thermal cycling and low-energy
impact on the fatigue behavior of carbon/PEEK laminates. Compos Part B Eng.
1999;30(8):849–65.

[186] Gómez-del Rı́o, T., Zaera, R., Barbero, E., Navarro, C. Damage in CFRPs due to low
velocity impact at low temperature. Compos Part B Eng. 2005;36(1):41–50.

[187] Vieille, B., Albouy, W., Bouscarrat, D., Taleb, L. High-temperature fatigue behaviour
of notched quasi-isotropic thermoplastic and thermoset laminates: Influence of matrix
ductility on damage mechanisms and stress distribution. Compos Struct.
2016;153:311–20.

[188] Wang, C.H., Gunnion, A.J. On the design methodology of scarf repairs to composite
laminates. Compos Sci Technol. 2008;68(1):35–46.

152
BIBLIOGRAPHY

[189] Rider, A.N., Wang, C.H., Chang, P. Bonded repairs for carbon/BMI composite at high
operating temperatures. Compos Part A Appl Sci Manuf. 2010;41(7):902–12.

[190] Archer, E., McIlhagger, A. Repair of damaged aerospace composite structures


[Internet]. Polymer Composites in the Aerospace Industry. Elsevier Ltd; 2015. 393–
412 p.

[191] Shufeng, L., Xiaoquan, C., Yunyan, X., Jianwen, B., Xin, G. Study on impact
performances of scarf-repaired carbon fiber reinforced polymer laminates. J Reinf
Plast Compos. 2015;34(1):60–71.

[192] Hou, Y., Tie, Y., Li, C., Sapanathan, T., Rachik, M. Low-velocity impact behaviors of
repaired CFRP laminates: Effect of impact location and external patch configurations.
Compos Part B Eng. 2019;163:669–80.

[193] Liu, H., Liu, J., Ding, Y., Zhou, J., Kong, X., Blackman, B.R.K., Kinloch, A.J.,
Falzon, B.G., Dear, J.P. Effects of Impactor Geometry on the Low-Velocity Impact
Behaviour of Fibre-Reinforced Composites: An Experimental and Theoretical
Investigation. Appl Compos Mater. 2020;

153
Appendices

Appendix A: Effects of impactor geometry on the low-


velocity impact behaviour of fibre-reinforced composites:
an experimental investigation

A.1 Introduction
The present appendix describes a detailed experimental investigation on the relatively
low-velocity (i.e. <10 m.s-1) impact behaviour of such composite laminates. In particular, the
effects of the geometry of the impactor have been studied and two types of impactor were
investigated: (a) a steel impactor with a hemispherical head and (b) a flat-ended steel impactor.
They were employed to strike the composite specimens with an impact energy level of 15 J.
After the impact experiments, all the composite laminates were inspected using ultrasonic C-
scan tests to assess the damage that was induced by the two different types of impactor.

A.2 Experimental programme

A.2.1 Impactors

Two types of impactors, a hemispherical steel (i.e. a round-nosed steel (RNS)) impactor
and a flat-ended (i.e. a flat-faced steel (FFS)) steel impactor, were employed in the experiments.
The hemispherically-headed steel impactor and the flat-ended steel impactor were
manufactured from stainless steel and are illustrated in Figure A-1. They had masses of 5.20
and 5.25 kg, respectively, and the impact velocities used were 2.40 m.s-1 and 2.39 m.s-1
respectively. The impact energy was 15 J in both cases. The impactor heads had a diameter of

154
APPENDIX A

16 mm, with the head of the flat-ended impactor having a chamfer of about 45o around its
periphery, as shown in Figure A-1. The diameter of the main body of the impactor was 20 mm
and the load cell, with a data sampling rate of 500 kHz, was located in the forward section of
the impactor.

(a) (b)
Figure A-1. Illustrations of: (a) the hemispherically-headed (RNS) steel and (b) the flat-
ended steel (FFS) impactor with a 45° chamfer on the edge to the flat-ended cylindrical
section.

A.2.2 Composite specimens

Composite laminates were manufactured from unidirectional (T700) carbon-fibre


reinforced epoxy-matrix pre-pregs. The stacked prepregs, provided by Beian Ltd, China, were
cured in an autoclave. Composite specimens for the drop-weight impact experiments were
produced from the manufactured composite panels and the cured composite laminates had a
thickness of 3 mm. All the composite specimens possessed the same lay-up of [+453/03/-
453/03]s. The dimensions of the composite specimens were 150 mm × 100 mm, as defined in
the ASTM D7136 standard [47].

A.2.3 Experimental procedures

The drop-weight impact experiments were performed using the CEAST drop-tower
system, as shown in Figure 3-5. In the experiments the hemispherically-headed impactor and
the flat-ended steel impactor were employed to strike the composite specimens with 15 J of
energy. Each target composite specimen was placed on a steel picture-frame which had outer

155
APPENDIX A

dimensions that matched those of the composite specimens and with a 125 mm × 75 mm cut-
out. This assembly was clamped to the base of the drop-weight tower using four toggle clamps
with rubber tips, which prevented slippage of the composite specimen during the impact test
[47]. For each test, three replicate composite specimens were tested and the signal from the
load cell was not subjected to any software filtering. All the tested composite specimens were
inspected using Sonatest C-scan device with 5 MHz probe, as shown in Figure 3-8, to assess
the extent of any impact-induced damage.

A.3 Experimental results

A.3.1 Loading response

The loading responses of the impacted composite specimens when struck by the
hemispherical steel and the flat-ended steel are shown in Figure A-2. At the same impact energy
level of 15 J, the average contact time of the flat-ended steel impacted specimens was about 1
ms (i.e. about 15%) shorter than that when using the hemispherical steel impactor. An average
maximum impact load of 7.1 kN was measured on the flat-ended steel impactor upon striking
the composite specimens. This is about 50% higher than that for the hemispherical steel
impactor, where the average maximum impact load was 4.7 kN. Now, the hemispherical steel
impactor provides a more localised loading of the composite specimen, relative to the flat-
ended impactor, and hence the load rises more slowly to the maximum load (i.e. 4.7 kN). The
maximum displacement measured for the composite specimen struck with the hemispherical
impactor was 6 mm. For the flat-ended impactor, the maximum load (i.e. 7.1 kN) is notably
higher and the maximum displacement in this case is reduced to 4.7 mm. The reasons for these
observations are that, as the flat-ended impactor has a larger area upon which to act on the
specimen, this impactor is arrested more quickly and the initial effective stiffness of the
specimen is higher.

156
APPENDIX A

(a) (b)
Figure A-2. Experimental results for (a) the load versus time and (b) the load versus
displacement curves obtained from the impact experiments performed using the
hemispherical steel (RNS) and the flat-ended steel (FFS) impactors. (Results from the
three replicate tests are shown.)

A.3.1 Impact damage

For both types of impactor it was found that, since the composite has a lay-up of
[+453/03/-453/03]s, there was a propensity for the delamination to grow at the interfaces between
the 45° and 0° plies, with the orientation of the lower ply determining the direction of
delamination propagation. Figure A-3 shows the delamination maps, measured via the C-scan
tests, obtained from the composite specimens impacted by the hemispherical steel and flat-
ended steel impactors. The right-hand side scale indicates the location of the delamination as a
function of the depth through the thickness of the specimen, where the red colour is
representative of the front (impacted) surface and the blue colour is representative of the rear
(non-impacted) surface of the composite specimen.)

The average areas of the damage maps produced by the hemispherical steel impactor
and flat-ended steel impactors are similar. However, the hemispherical steel impacted
specimens show a continuous damage area, which is centred around the point of impact, whilst
the specimens impacted with the flat-ended steel impactor exhibit two separate damage areas,
with a central zone exhibiting no damage. These differences in the extent and shape of the
damage are explained below from the results of the numerical modelling studies.

157
APPENDIX A

Figure A-3. Interlaminar C-scan damage maps obtained from the specimens impacted with
the hemispherical steel (RNS) and the flat-ended steel (FFS) impactors. (Results from the
three replicate tests are shown. The 0° fibre direction is indicated.)

A.4 Summary
A hemispherical steel impactor and a flat-ended steel impactor have been used to impact
carbon fibre-reinforced/epoxy-matrix composite laminates with an energy level of 15 J. All the
composite specimens were then inspected using ultrasonic C-scan testing to assess the impact
damage. It was found that the composite specimens struck with a flat-ended steel impactor
suffered a higher maximum load upon impact, but experienced a smaller out-of-plane
displacement, than those struck with the hemispherical steel impactor. The comparison of the
experimental C-scan images showed that the damage maps of the composite specimens
impacted with the hemispherical steel and the flat-ended steel impactors were similar in area
but with somewhat more areal damage occurring when the hemispherical steel impactor was
used. However, the composite specimens impacted with the hemispherical steel impactor
showed a continuous damage area, which was centred around the point of impact, whilst the
flat-ended steel impacted specimens exhibited two separate damage areas, with a central zone

158
APPENDIX A

exhibiting no damage. These studies have increased our understanding of the behaviour of
composite laminates when subjected to a relatively low-velocity impact and have proposed,
and validated, an elastic-plastic three-dimensional numerical model for predicting the response
of the laminate [193].

159
Appendix B: Summary of copyright permissions

Page Type Source Copyright holder Permission requested Granted?


8 Figure 2-1 Strong [4] © 2008 Society of Manufacturing Engineers 12/09/2020 Yes
9 Figure 2-2 Mallick [3] © 2007 Taylor & Francis Group, LLC 11/09/2020 Yes
11 Figure 2-3 Tong et al. [25] © 2002 Elsevier Science Ltd. 10/09/2020 Yes
11 Figure 2-4 Bergstorm [28] © 2015 Elsevier Inc. 10/09/2020 Yes
14 Figure 2-5 Strong [4] © 2008 Society of Manufacturing Engineers 12/09/2020 Yes
15 Figure 2-6 Mallick [3] © 2007 Taylor & Francis Group, LLC 11/09/2020 Yes
17 Figure 2-7 Cantwell and Morton [46] © 1989 Butterworth & Co. Ltd. 10/09/2020 Yes
18 Figure 2-8 ASTM International [47] © ASTM International 10/09/2020 Yes
21 Figure 2-9 Will et al. [55] © 2002 Elsevier Science Ltd. 10/09/2020 Yes
24 Figure 2-10 Davies and Olsson [43] © Royal Aeronautical Society 2004 10/09/2020 Yes
25 Figure 2-11 Pagano [81] © 1989 North Holland 11/09/2020 Yes
25 Figure 2-12 Laval [90] © 2003 Elsevier Science Ltd. 10/09/2020 Yes
26 Figure 2-13 Daniel et al. [91] © 2007 Elsevier Ltd. 10/09/2020 Yes
27 Figure 2-14 Van Paepegem et al. [97] © 2005 Elsevier Ltd. 10/09/2020 Yes
28 Figure 2-15 Matthews [22] © 1989 Elsevier Science & Technology 11/09/2020 Yes
33 Figure 2-16 Puck and Schürmann [125] © 2002 Elsevier Science Ltd. 10/09/2020 Yes
34 Figure 2-17 Faggiani and Falzon [135] © 2010 Elsevier Ltd. 10/09/2020 Yes

160

You might also like