You are on page 1of 287

Fabricating of Multiscale Composite Materials

Based on TPU reinforced by carbon fibre and


Graphene Nanoplatelets (GNPs)

A thesis submitted to the University of Manchester for the degree of

Doctor of Philosophy

in the Faculty of Science and Engineering

2019

Haneen Zuhair Naji Naji

School of Materials

University of Manchester
Table of contents

Table of contents ......................................................................................................... 2

List of Figures ............................................................................................................. 8

List of Tables ............................................................................................................. 17

List of Abbreviations ................................................................................................. 20

List of Symbols .......................................................................................................... 23

Thesis Abstract .......................................................................................................... 25

Declaration ................................................................................................................ 26

Copyright Statement .................................................................................................. 27

Dedication .................................................................................................................. 28

Acknowledgment ....................................................................................................... 29

Chapter 1: Introduction ........................................................................................... 30

1.1 Background............................................................................................... 30

1.2 Global Market of CF composites ............................................................ 32

1.3 Research Aims and Objectives ................................................................ 34

1.4 Thesis Structure ....................................................................................... 35

Chapter 2: Literature Review ................................................................................... 37

2.1 Thermoplastic and Thermoplastic Elastomers ..................................... 37


2.1.1 Thermoplastic Polyurethane...................................................................... 38
2.1.1.1 Morphology of TPU Structure ....................................................... 42
A) Effect of weight percentage of Hard Segments .................................... 45
B) Effect of Annealing Treatment on the Morphology of TPU structure .. 47

2.2 Nanocomposites and Nanocomposite Matrices ..................................... 51


2.2.1 Graphitic Forms of Carbon ....................................................................... 51
2.2.2 Graphene Nanoplatelets (GNPs) ............................................................... 52
2.2.3 Processing Methods of Polymer/GNP Nanocomposites ........................... 54
2.2.3.1 In situ Polymerisation .................................................................... 54
2.2.3.2 Solution Mixing ............................................................................. 55
2
2.2.3.3 Melt Compounding ........................................................................ 55
2.2.4 Properties of GNP Nanocomposites ......................................................... 57
2.2.4.1 Dispersion and Melt Rheology of GNP Nanocomposites ............. 57
2.2.4.2 Mechanical, Electrical and Thermal Properties of Nanocomposites
60

2.3 Summary ................................................................................................... 62

2.4 Multiscale Composites ............................................................................. 63


2.4.1 Introduction to Multiscale Composites ..................................................... 63
2.4.2 Carbon-Fibre Reinforcement .................................................................... 64
2.4.3 Manufacturing of Multiscale Composites ................................................. 66
2.4.4 Effect of Nanofillers on the Properties of MSCs ...................................... 66
2.4.4.1 Tensile and Flexural Properties ...................................................... 67
2.4.4.2 Interlaminar Shear Strength (ILSS) ............................................... 68
2.4.4.3 Interlaminar Fracture Toughness (ILFT) ....................................... 70
2.4.4.4 Impact-damage Resistance and Damage Tolerance ....................... 72
2.4.4.5 Electrical and Thermal Properties .................................................. 76

2.5 Summary ................................................................................................... 78

Chapter 3: Experimental Methods .......................................................................... 80

3.1 Introduction .............................................................................................. 80

3.2 Materials ................................................................................................... 80


3.2.1 Raw Materials of TPU-70 HS Synthesis................................................... 80
3.2.1.1 Isocyanate....................................................................................... 80
3.2.1.2 Polyol ............................................................................................. 81
3.2.1.3 Chain Extender ............................................................................... 81
3.2.1.4 Catalyst........................................................................................... 82
3.2.1.5 Solvent ........................................................................................... 82
3.2.2 Conductive Raw Materials ........................................................................ 83
3.2.2.1 Graphene Nanoplatelets (GNPs) .................................................... 83
3.2.2.2 Carbon Fibre .................................................................................. 83

3.3 Synthesis of Unfilled TPU........................................................................ 84


3.3.1 Preparation of Synthesis Apparatus .......................................................... 84
3.3.2 Preparation of Chemical Substance for Synthesis .................................... 84
3
3.3.3 Chemistry of Synthesis of Neat TPU-70 HS............................................. 85
3.3.4 Calculations for Formulation of TPU ....................................................... 87

3.4 Synthesis of TPU70-NCs using In situ Polymerisation ......................... 92


3.4.1 Solvent-casting Films ................................................................................ 93

3.5 Manufacturing Process ............................................................................ 94


3.5.1 Injection-moulding (IM) Process .............................................................. 94
3.5.2 Compression-moulding (CM) Process ...................................................... 94
3.5.2.1 Compression-moulding for Preparing Samples of Neat TPU-70 HS
and TPU70-NCs ............................................................................................. 95
3.5.2.2 Preparation of Multiscale Composites (MSCs).............................. 96

3.6 Cutting Process for Preparing TPU70/CF Laminate and MSCs


Samples ................................................................................................................. 97

3.7 Annealing Treatment ............................................................................... 97

3.8 Characterisation and Experimental Procedure .................................... 98


3.8.1 Gel Permeation Chromatography (GPC) .................................................. 98
3.8.2 Differential Scanning Calorimetry (DSC) .............................................. 100
3.8.3 Dynamic Mechanical Thermal Analysis (DMTA) ................................. 103
3.8.4 Thermogravimetric Analysis (TGA) ....................................................... 106
3.8.5 Thermal Conductivity ............................................................................. 107
3.8.6 Wide-Angle X-ray Diffraction (WAXD) ................................................ 108
3.8.7 X-ray Photoelectron Spectroscopy (XPS) ............................................... 110
3.8.8 Elemental Analysis (CHNS Analysis) .................................................... 111
3.8.9 Scanning Electron Microscopy (SEM) ................................................... 111
3.8.10 Transmission Electron Microscopy (TEM) .......................................... 114
3.8.11 Optical Microscopy ............................................................................... 115

3.9 Rheological Characterisation ................................................................ 115


3.9.1 Background of Rheology ........................................................................ 115
3.9.2 Experimental Procedure of Rheological Test ......................................... 117

3.10 Impedance Spectroscopy ....................................................................... 117

3.11 Infrared (IR) Thermography ................................................................ 119

3.12 Analysis of Ultrasonic C-Scan Technique ............................................ 120

4
3.13 Acid Digestion Method .......................................................................... 121

3.14 Mechanical Testing ................................................................................ 123


3.14.1 Tensile Test ........................................................................................... 123
3.14.1.1 Tensile Test of Matrices ............................................................... 123
3.14.1.2 Tensile Test of TPU70/CF and Multiscale Composites ............... 124
3.14.2 Flexural Test.......................................................................................... 125
3.14.3 Interlaminar Shear Strength (ILSS) ...................................................... 126
3.14.4 Mode-I Interlaminar Fracture Toughness ............................................. 127
3.14.5 Impact Resistance Test .......................................................................... 129
3.14.6 Compression after Impact (CAI) ........................................................... 130

Chapter 4: Effect of Graphene Nanoplatelet Size and Annealing Treatment on the


Properties of a Matrix of TPU-70 Hard Segments ................................................ 132

4.1 Introduction ............................................................................................ 132

4.2 Characterisation of Graphene Nanoplatelets ...................................... 133


4.2.1 Thermal Stability of GNPs ...................................................................... 133
4.2.2 Characterisation of GNP Morphology .................................................... 133
4.2.2.1 TEM and SEM Analysis .............................................................. 133
4.2.2.2 WAXD Analysis .......................................................................... 136

4.3 TPU-70 HS NCs Characterisation ........................................................ 137


4.3.1 CHNS and XPS analysis ......................................................................... 137
4.3.2 Gel Permeation Chromatography (GPC) ................................................ 140
4.3.3 Structure and Morphology Characterisation ........................................... 142
4.3.3.1 SEM Analysis .............................................................................. 142
4.3.3.2 Optical Microscopic (OM) Analysis ............................................ 144
4.3.3.3 WAXD Analysis .......................................................................... 145
4.3.4 Thermal Analysis .................................................................................... 147
4.3.4.1 TGA ............................................................................................. 147
4.3.4.2 DSC Analysis ............................................................................... 150
4.3.4.3 Dynamic Mechanical Thermal Analysis (DMTA) ...................... 155
4.3.5 Melt Rheology of TPU70/GNPs Nanocomposites ................................. 160
4.3.5.1 Oscillatory Strain Sweep .............................................................. 160
4.3.5.2 Frequency Sweep ......................................................................... 162

5
4.3.6 Mechanical Properties ............................................................................. 166
4.3.6.1 Tensile Test of TPU-70 HS and TPU70/GNPs-NCs ................... 166
4.3.6.2 Flexural Properties of TPU-70 HS and TPU70-NCs ................... 171
4.3.7 Testing Conductive Properties ................................................................ 174
4.3.7.1 Electrical Conductivity................................................................. 174
4.3.7.2 Thermal Conductivity .................................................................. 176

4.4 Summary ................................................................................................. 178

Chapter 5: Results and Discussion of TPU70/GNPs/CF Multiscale Composites


(MSCs) ..................................................................................................................... 181

5.1 Introduction ............................................................................................ 181

5.2 Void Content and Fibre Volume Fraction ........................................... 182

5.3 DMTA Analysis ...................................................................................... 183

5.4 Mechanical Testing ................................................................................ 189


5.4.1 Tensile Test ............................................................................................. 189
5.4.1.1 Theoretical Micromechanical Analysis Model ............................ 194
5.4.2 Flexural Test............................................................................................ 197
5.4.3 Interlaminar Shear Strength (ILSS) ........................................................ 201

5.5 Electrical Conductivity .......................................................................... 206

5.6 Thermal Conductivity............................................................................ 211

5.7 Summary ................................................................................................. 215

Chapter 6: The Effect of GNPM25 on the Delamination and Damage Tolerance


of TPU70/CF Composites ....................................................................................... 217

6.1 Introduction ............................................................................................ 217

6.2 Micromechanical Modelling Analysis .................................................. 218


6.2.1 Halpin-Tsai Model .................................................................................. 218
6.2.1.1 Effect of Variation in Filler Modulus and Aspect Ratio .............. 220
6.2.1.2 Prediction of Tensile Modulus of TPU70/GNP Nanocomposites 223

6.3 Mechanical Testing of Multiscale Composites .................................... 225


6.3.1 Mode-I Interlaminar Fracture Toughness Tests ...................................... 225
6.3.2 Impact Damage Resistance ..................................................................... 236

6
6.3.2.1 Impact Damage Evaluation using SEM ....................................... 247
6.3.3 Compression after Impact (CAI) ............................................................. 249

6.4 Summary ................................................................................................. 253

Chapter 7: Conclusions and Suggestions for Future Work ................................. 256

7.1 Conclusions ............................................................................................. 256


7.1.1 Introduction ............................................................................................. 256
7.1.2 Characterisation of Graphene Nanoplatelets ........................................... 256
7.1.3 Characterisation of TPU-70 HS Nanocomposites .................................. 257
7.1.4 Characterisation of Multiscale Composites ............................................ 259
7.1.5 Damage Tolerance and Delamination Resistance of TPU70/CF and MSCs
.......................................................................................................................... 260

7.2 Suggestions for Future Work ................................................................ 260

References................................................................................................................ 262

Appendix A .............................................................................................................. 284

Appendix B .............................................................................................................. 287

Words count: 67,509

7
List of Figures

Figure 1.1: The market of carbon fibre composites by the end of 2016 and 2024 in
United States estimated in millions dollars [31] ........................................................ 33
Figure 1.2: The revenue of matrices of carbon fibre composites in 2013 in United
States regions[32] ....................................................................................................... 33
Figure 2.1: Schematic representation of TPU structure: (A) repeat hard and soft
segment units, and (B) alternating structure of hard segments (HS) and soft segments
(SS) [45] ..................................................................................................................... 38
Figure 2.2: Urethane linkage [60] .............................................................................. 40
Figure 2.3: Schematic representation depicting four types of H-bonds relating to
polyether-based polyurethane, where (I), (II) and (III) are urethane–urethane
interactions, while (IV) is a urethane-soft segment interaction [66, 67] .................... 41
Figure 2.4: Schematic representation of the microphase morphology of TPU structure
[63] ............................................................................................................................. 42
Figure 2.5: Thermal transition of segmented PU by using DSC techniques [70] ...... 43
Figure 2.6: Schematic representations of the morphological changes of segmented
polyurethane elastomers obtained through DSC scans (a) below the microphase
mixing temperature (TII), (b) between the melting temperature and TII, and (c) above
the temperature of the microcrystalline hard domain (TIII) [74] ................................ 44
Figure 2.7: Comparison of tensile test results (modulus, yield strength and elongation
at break) of TPU based on the different hard segments: a) TPU with 60 wt.% HS; b)
TPU with 70 wt.% HS; and c) TPU with 80 wt.% HS before and after annealing [46]
.................................................................................................................................... 46
Figure 2.8: DSC results for PU-65%HS at 120 °C at different annealing times (ta)
obtained at a heating rate of 20 °C/min [71] .............................................................. 47
Figure 2.9: Schematic representation of a morphological model illustrating the TPU
systems for different phases. (a) Melt-quenched samples, (b) microphase separated
samples, (c) microphase separated samples under mechanical stress [73] ................ 48
Figure 2.10: Schematic representation of the development of an HS-rich domain in a
TPU structure at different annealing temperatures (Ta): (a) sample directly after
melt-quenching (MQ), b) MQ sample annealed at 25°C, and c) MQ samples
annealed at higher Ta [82].......................................................................................... 50

8
Figure 2.11: Graphite structure illustrating the sp2 hybridised carbon atoms bonded in
continuous hexagons [85] .......................................................................................... 51
Figure 2.12: Graphene as the mother of all graphitic forms: when stacked, it forms
3D graphite, or can be wrapped into 0D fullerenes and rolled into 1D nanotubes [86,
88] .............................................................................................................................. 52
Figure 2.13: Schematic diagram of the modifications to graphite structure for GNP
preparation [22] .......................................................................................................... 53
Figure 2.14: Comparison of thermal conductivity of TPU70/GNP nanocomposites
made by in-situ polymerisation, solution mixing and melt compounding [46] ........ 56
Figure 2.15: Schematic representation of three common descriptions of the
dispersion qualities of layered nanofillers within polymer matrices [97] .................. 58
Figure 2.16: Dynamic frequency sweeps of PC/G melts after annealing for (a) 10,000
and (b) 20,000 s, and PC/FGS melts after annealing for (c) 10,000 and (d) 20,000 s at
230 °C [106] ............................................................................................................... 59
Figure 2.17: Young's modulus as a function of crystallinity of PLA (XC (%)) for
PLA/GNP-S and PLA/GNP-L, showing a robust relation between modulus and
crystallinity for the system based on GNP-L [108] ................................................... 61
Figure 2.18: SEM images to describe the failure modes of (a) controlled epoxy/CF
composites and (b) epoxy/CF incorporating 1 wt.% GNPs [131] ............................. 68
Figure 2.19: Effect of NH2-MWCNT concentration on the ILSS values of epoxy/GF
composites [133] ........................................................................................................ 69
Figure 2.20: Schematic representation of three loading modes during interlaminar
fracture toughness test [136] ...................................................................................... 70
Figure 2.21: SEM images of failure mode after mode-I ILFT test of epoxy/0.5 wt.%
SCCNFs/CF laminates. (a) Peel off and pull-out of SCCNFs, and b) unravelling of
SCCNFs [137] ............................................................................................................ 71
Figure 2.22: Schematic representing typical damage modes of low-energy impact of
laminated composites [6] ........................................................................................... 73
Figure 2.23: The effect of incorporating MWCNTs within epoxy/CF composites on
the a) delamination area against impact energy levels and b) the percentage of
absorbed energy against energy levels for MSCs and neat composites [122] ........... 75
Figure 3.1: Chemical structure of 4,4’-methylenebis phenyl isocyanate [149] ......... 81
Figure 3.2: Chemical structure of PEG-PPG-PEG polyol [149]................................ 81
Figure 3.3: Chemical structure of chain extender (1,5-pentanediol) [149] ................ 81
9
Figure 3.4: Chemical structure of 1,4-Diazabicyclo[2.2.2]octane [149] ................... 82
Figure 3.5: Chemical structure of N,N-Dimethylacetamide [149] ........................... 82
Figure 3.6: Chemical reaction of TPU synthesis. A) Step one: synthesis of pre-
polymer and B) step two: synthesis of TPU-70 HS ................................................... 86
Figure 3.7: Synthesis process set-up during first and second steps of TPU-70 HS
polymerisation process ............................................................................................... 87
Figure 3.8: Dispersion of GNPs in DMAc solvent using an ultra-sonicated water bath
technique: A) before sonication and B) after sonication ............................................ 92
Figure 3.9: Solvent-casting of A) TPU-70 HS and B) TPU70-NCs films in PTFE
moulds ........................................................................................................................ 93
Figure 3.10: A) Haake Minijet II injection-moulding machine, B) dumbbell-shaped
mould, and C) some of the resultant nanocomposite samples ................................... 94
Figure 3.11: Schematic representation of the cyclic hot-press consolidation process
used for manufacturing neat TPU-70 HS and TPU70NCs samples .......................... 96
Figure 3.12: Schematic representation of the compression-moulding process for
TPU70/CF laminate and multiscale composite: A) preheated mould, B) arrangement
of TPU films and CF layers, C) resultant composite ................................................. 97
Figure 3.13: Influence of molecular weight on the mechanical strength of a polymer
[156, 157] ................................................................................................................... 98
Figure 3.14: Separation process in GPC column based on disparity in molecular size
[70, 159] ................................................................................................................... 100
Figure 3.15: Design of DSC instruments: a) heat flux DSC, and b) power
compensation DSC; A) furnace, B) separate heater and C) samples and reference
holders [162] ............................................................................................................ 101
Figure 3.16: The relation between stress and strain during dynamic mechanical
testing [70, 165, 166] ............................................................................................... 104
Figure 3.17: Argand diagram to explain the relation between complex modulus and
its components [70, 165, 166] .................................................................................. 105
Figure 3.18: A) TGA curve, and B) DGA curve [70, 169] ...................................... 107
Figure 3.19: Schematic representation of the X-ray diffractions from the planes of a
crystal lattice found in an ordered system [52, 172] ................................................ 109
Figure 3.20: Schematic illustration of beam-sample interaction shows backscattered
electrons (BSE), secondary electrons (SE) and X-rays [176] .................................. 112

10
Figure 3.21: A) coating machine set up, B) carbon fibre braid for coating neat TPU-
70 HS and TPU70-NCs and C) the holder of sample with the resultant coated sample,
D) Pu/Pd alloy film for coating TPU70/CF laminate and MSCs and E) the sample
holder with the resultant coated sample. .................................................................. 114
Figure 3.22: Schematic representation of Newtonian fluid for a parallel plate model
[1, 181] ..................................................................................................................... 116
Figure 3.23: Some electrical conductivity samples that were used during the present
work: A) in-plane direction and B) out-of-plane direction ...................................... 118
Figure 3.24: Set-up of infrared thermography experiment ...................................... 119
Figure 3.25: Schematic of through-transmission ultrasonic C-scan [191] ............... 121
Figure 3.26: Tensile specimen (dumbbell-shaped specimen)[193] ......................... 123
Figure 3.27: Set-up of a specimen in the tensile test machine with video
extensometer ............................................................................................................ 124
Figure 3.28: Flexural test set-up for A) TPU-70 HS nanocomposites and B)
multiscale composites .............................................................................................. 125
Figure 3.29: Interlaminar shear strength (ILSS) failure modes: 1) interlaminar shear,
2) tension or compression failure, and 3) inelastic deformation [200] .................... 127
Figure 3.30: Double cantilever beam (DCB) specimen geometry [135] ................. 127
Figure 3.31: DCB specimen of mode I interlaminar fracture toughness test specimen
.................................................................................................................................. 128
Figure 3.32: schematic representative the calculation of  in MBT ........................ 129
Figure 3.33: Schematic representation of impact test set-up [204] .......................... 130
Figure 3.34: Schematic representation of compression after impact (CAI) fixture
[207] ......................................................................................................................... 130
Figure 4.1: TGA curve of as-received graphite, GNPM5, GNPM15 and GNPM25
with average particle diameter 5m, 15 m and 25m respectively under a nitrogen
atmosphere ............................................................................................................... 133
Figure 4.2: SEM images of GNPs before sonication A) GNPM5 (5m average
particle diameter), C) GNPM15 (15m average particle diameter) and E) GNPM25
(25m average particle diameter) and after sonication B) GNPM5, D) GNPM15 and
F) GNPM25. The arrows indicate the folded and rolled up GNPs .......................... 134
Figure 4.3: TEM images of GNPs before sonication: A) GNPM5, C) GNPM15 and
E) GNPM25; and after sonication: B) GNPM5, D) GNPM15 and F) GNPM25. The

11
arrows in the images indicate the clear wrinkling and folding areas of the different
GNP types ................................................................................................................ 135
Figure 4.4: WAXD for as-received GNPs (M5, M15 and M25), donated by A-GNPs
(M5, M15 and M25) before sonication and compared to WAXD for S-GNPs (M5,
M15 and M25) after sonication ................................................................................ 136
Figure 4.5: Low-resolution XPS survey scan of GNP types (M5, M15, and M25): A)
as received; B) after first wash; C) after second wash ............................................. 138
Figure 4.6: 𝑀𝑛 and 𝑀𝑤 for neat TPU-70 HS and extracted TPU-70 HS matrix from
its nanocomposites obtained from Gel permeation chromatograph test. ................. 141
Figure 4.7: SEM images of cryogenic-fracture surfaces of A) neat TPU-70 HS, B)
TPU70-NCs incorporated with GNPM5, C) GNPM15, and D) GNPM25 at low and
high magnifications. The yellow rectangular area in the low magnified images on the
left-hand side indicates the magnified region on the right-hand side. The yellow
arrows in the highly magnified images indicate the GNPs flakes, while the red curves
indicate the path of electrons due to form conductive network of GNPM15 and
GNPM25 .................................................................................................................. 143
Figure 4.8: Optical micrograph images of the TPU70/GNPs-NCs based on different
GNP sizes: A) GNPM5, B) GNPM15, and C) GNPM25, respectively .................. 144
Figure 4.9: WAXD patterns of neat TPU-70 HS and TPU70-NCs incorporated with
5wt.% of GNPM5, GNPM15 and GNPM25 A) before annealing and B) after
annealing .................................................................................................................. 146
Figure 4.10: A) TGA and B) DTG thermograms for TPU-70 HS and TPU70/GNP
nanocomposites based on various GNP sizes .......................................................... 148
Figure 4.11: DSC curves for TPU-70 HS and TPU70-NCs incorporated with 5 wt.%
of GNPM5, GNPM15 and GNPM25 with average particle diameter 5, 15 and 25 m
respectively at 10° C/min heating/cooling rate in order to show the effect of GNP
sizes and annealing treatment on their melting temperature, glass transition
temperature and crystalline temperature. Where A), C) and E) are DSC scan cycles
before annealing and B), D) and F) are DSC scan cycles after annealing ............... 151
Figure 4.12: Log storage modulus, E’, for neat TPU-70 HS and TPU70/GNPs/NCs
based on different GNP sizes (5, 15 and 25m denoted by M5, M15 and M25
respectively) versus temperature (°C): A) before annealing (Non- annealed) and B)

after annealing (Annealed) ....................................................................................... 155

12
Figure 4.13: Comparison of storage modulus of neat TPU-70 HS and TPU70/GNPs-
NCs based on the different sizes of GNPs, before and after annealing at three
different temperatures (-100 °C, 25 °C, and 100 °C, respectively) ......................... 156
Figure 4.14: Tan for neat TPU-70 HS and TPU70/GNPs-NCs based on different
GNP sizes (5, 15 and 25 denoted by M5, M15 and M25) A) before annealing and B)
after annealing .......................................................................................................... 158
Figure 4.15: Oscillatory shear measurement results at constant frequency 1Hz for
TPU-70 HS and TPU70/GNPs-NCs at 200 C show the independence of G’ and G”
versus strain amplitude ............................................................................................. 161
Figure 4.16: Frequency sweep measurements conducted within LVR ( = 0.2 %) at
200C show the values of A) Shear storage modulus, G’ and B) shear loss modulus,
against frequency for TPU-70 HS and TPU70-NCs ................................................ 163
Figure 4.17: Complex viscosity, *, of neat TPU-70 HS and Its NCs frequency
sweep measurements conducted within LVR ( = 0.2%) show the influence of
incorporating with different sizes of GNPs at same loading 5 wt. %. ..................... 165
Figure 4.18: Stress-strain curves for TPU-70 HS and TPU70/GNPs-NCs: A, B)
before annealing and C, D) after annealing. The right hand curves B and D represent
a zoomed-out view of the elastic region of the initial part of the tensile stress-strain
curve A and C respectively ...................................................................................... 167
Figure 4.19: Tensile properties of neat TPU-70 HS and TPU70/GNPs-NCs based on
different sizes of GNPs before and after annealing ................................................. 168
Figure 4.20: Typical flexural stress-strain curve for neat TPU-70 HS and
TPU70/GNPs-NCs based on various size of GNPs (5,15 and 25 m denoted by M5,
M15 and M25 respectively): A) before annealing (Non-annealed) and B) after
annealing (Annealed) ............................................................................................... 171
Figure 4.21: Typical sample of Neat TPU-70 HS and TPU70/GNPs-NCs after
subjecting to flexural test shows a type of flexural failure ...................................... 171
Figure 4.22: Flexural properties of neat TPU-70 HS and TPU70/GNPs-NCs based
on various size of GNPs (5,15 and 25 m denoted by M5, M15 and M25
respectively) before and after annealing .................................................................. 172
Figure 4.23: Electrical conductivity versus frequency for TPU-70 HS and
TPU70/GNPs-NCs in two directions: A) In-plane direction and B) Out-of-plane
direction.................................................................................................................... 175

13
Figure 4.24: Thermal conductivity of neat TPU-70 HS compared to TPU70/GNPs-
NCs based on the various size of GNPs ................................................................... 177
Figure 5.1: Storage modulus of A) non-annealed and B) annealed TPU70/CF
composites and MSCs incorporated with 5wt.% of GNPs based on their various sizes
of GNPs as a function of temperature ...................................................................... 184
Figure 5.2: Tan delta (loss factor) of TPU70/CF laminate and multiscale composites
A) before annealing and B) after annealing ............................................................. 187
Figure 5.3: Typical tensile stress-strain curves of TPU70/CF and TPU70/5 wt.
%GNPM5, TPU70/5 wt.%GNPM15 and TPU70/5 wt.%GNPM25/CF-MSCs A)
before annealing and B) after annealing .................................................................. 189
Figure 5.4: A) Ultimate tensile strength and elastic modulus and B) elongation at
break % of TPU70/CF and TPU70/GNPs/CF-MSCs before and after annealing ... 191
Figure 5.5: Failure pattern of A) TPU70/CF laminates, B) TPU70/GNPM5/CF, C)
TPU70/GNPM15/CF and D) TPU70/GNPM25/CF multiscale composites after
tensile test ................................................................................................................. 193
Figure 5.6: Comparison between predicted elastic modulus and results of
experimental measurement of TPU70/CF and MSCs as functions of different
matrices .................................................................................................................... 196
Figure 5.7: Typical flexural stress-strain curves of TPU70/CF and MSCs A) before
and B) after annealing .............................................................................................. 198
Figure 5.8: Normalised flexural strength and flexural modulus for TPU70/CF and
MSCs before and after annealing the normalisation have been done according to the
high Vf percentage (Vf=52%) .................................................................................. 199
Figure 5.9: Stress-strain curves after short-beam shear test of TPU70/CF laminates
and TPU70/5 wt.%GNPs/CF multiscale composite A) before annealing and B) after
annealing .................................................................................................................. 201
Figure 5.10: Normalised ILSS of TPU70/CF and MSCs according to void-free
composite equations and Vf = 47% before and after annealing .............................. 203
Figure 5.11: SEM micrographs of A) TPU70/CF composites and B) TPU70/5
wt.%GNPM5/CF/MSCs after a short beam shear test taken from a side view of the
samples. The right-hand images represent magnified images of the positions circled
in yellow positions in the left-hand image ............................................................... 205
Figure 5.12: SEM micrographs of C) TPU70/5 wt.%GNPM15/CF/MSCs and D)
TPU70/5 wt.%GNPM25/CF/MSCs after a short-beam shear test taken from a side
14
view of the samples. The right-hand images represent magnified images of the
positions circled in yellow positions in the left-hand image .................................... 206
Figure 5.13: A) In-plane and B) out-of-plane electrical conductivity of TPU70/CF
laminates and MSCs incorporating 5 wt.% GNPs of various size (5, 15 and 25 m
denoted by M5, M15 and M25 respectively) ........................................................... 208
Figure 5.14: Pixel versus thermal diffusivity () on the right side and mapping of
thermal diffusivity on the left side of A) TPU70/CF, B) TPU70/5 wt.%GNPM5/CF,
C) TPU70/5 wt.%GNPM15/CF and D) TPU70/5 wt.%GNPM25/CF composites.. 212
Figure 6.1: Halpin-Tsai model showing the effect of Ef on the tensile modulus of
TPU-70 HS nanocomposites .................................................................................... 221
Figure 6.2: Calculation of Halpin-Tsai model showing the effect of loading and
aspect ratio of GNPs on the tensile modulus of TPU70-NCs incorporated with 0.5
wt.%, 2.5 wt.% and 5wt.% GNPM25. The letters U and L indicate the upper and
lower bound respectively ......................................................................................... 222
Figure 6.3: Tensile modulus of elasticity of TPU70/GNPM25/NCs: Comparison
between Halpin-Tsai model, 2D MH-T and 3D MH-T models with experimental data
.................................................................................................................................. 223
Figure 6.4: Representation of cube-root compliance C1/3 versus crack length, a, of
A) TPU70/CF; B) MSCs incorporated with 0.5 wt.% GNPs; C) MSCs incorporated
with 2.5 wt.% GNPs; and D) MSCs incorporated with 5 wt.% GNPs .................. 226
Figure 6.5: Typical load-displacement curves of Mode-I test (DCB) of TPU70/CF
laminates and MSCs incorporated with 0.5, 2.5, and 5 wt. % GNPs with 25 m
average particle sizes................................................................................................ 227
Figure 6.6: Typical R-curves (critical strain energy release rate, GIC) of TPU70/CF
and MSCs versus crack length ................................................................................. 229
Figure 6.7: SEM images of TPU70/CF composites at fracture surface in Mode-I
fracture toughness testing A) near the inserted film and B) far from the inserted film.
The yellow lines represent the hackles, while the blue lines represent the river
patterns. The yellow rectangular area indicates the magnified region ..................... 231
Figure 6.8: SEM images of MSCs combined with 0.5 wt.% GNPs at fracture surface
of Mode-I fracture toughness testing A) near the inserted film and B) far from the
inserted film. The yellow lines represent the hackles, while the blue lines represent
the river patterns. The yellow rectangular area indicates the magnified region. ..... 232

15
Figure 6.9: SEM images of MSCs combined with 2.5 wt.% GNPs at fracture surface
of Mode-I fracture toughness testing A) near the inserted film and B) far from the
inserted film. The red lines represent the agglomeration regions of GNPs ............. 233
Figure 6.10: SEM images of MSCs incorporated with 5 wt.% GNPs at fracture
surface of Mode-I fracture toughness testing A) near the inserted film and B) far
from the inserted film. The red lines represent the agglomeration regions of GNPs
.................................................................................................................................. 234
Figure 6.11: The response of contact force versus time of TPU70/CF laminate and
multiscale composite incorporated with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at
various impact energy levels: A) 5 J, B) 15 J and C) 25 J, and D) the comparison of
peak force at these impact energy levels .................................................................. 237
Figure 6.12: The corresponding contact force versus displacement of TPU70/CF
laminates and multiscale composites combined with 0.5 wt.%, 2.5 wt.% and 5 wt.%
GNPs at different impact energy levels: A) 5 J, B) 15 J and C) 25 J, and D) the
comparison of peak deformation at these impact energy levels .............................. 240
Figure 6.13: The response of energy versus time of TPU70/CF laminate and
multiscale composite combined with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at various
impact energy levels: A) 5 J, B) 15 J and C) 25 J, and D) the comparison of absorbed
energy at these impact energy levels ........................................................................ 242
Figure 6.14: Damage area measurements of TPU70/CF and MSCs combined with 0.5
wt.%, 2.5 wt.% and 5 wt.% GNPs at various impact energy levels ......................... 244
Figure 6.15: Comparison of dent depth recovery measurements of TPU70/CF
laminates and MSCs incorporated with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at
impact energy levels of A) 5 J, B) 15 J and C) 25 J respectively ............................ 246
Figure 6.16: SEM cross-section images after subjection to 25 J impact energy of A)
TPU70/CF laminates and B) MSCs incorporated with 0.5 wt.% GNPM25 , C) MSCs
incorporated with 2.5 wt.% GNPM25 and D) MSCs incorporated with 5 wt.%
GNPM25 .................................................................................................................. 249
Figure 6.17: Compression stress versus extension curves of TPU70/CF laminates and
MSCs combined with 0.5 wt. %, 2.5 wt. % and 5 wt. % GNPs after various impact
energy levels: A) 0 J, B) 5 J, C) 15 J and D) 25 J, respectively ............................... 250
Figure 6.18: A) Ultimate compressive and CAI strength and B) residual compressive
strength (%) of TPU70/CF and MSCs combined with 0.5 wt.%, 2.5 wt.% and 5 wt.%
GNPs with varying impact energies ........................................................................ 252
16
List of Tables

Table 2.1: Carbon fibre properties [120] ................................................................... 65


Table 3.1: Properties of 4,4’-methylenebis phenyl isocyanate [149]......................... 80
Table 3.2: Properties of PEG-PPG-PEG polyol [149] ............................................... 81
Table 3.3: Properties of chain extender (1,5-pentanediol) [149] ............................... 82
Table 3.4: Properties of 1,4-Diazabicyclo[2.2.2]octane [149] ................................... 82
Table 3.5: Properties of N,N-Dimethylacetamide [149] ............................................ 83
Table 3.6: Typical properties of grade-M GNPs [150] .............................................. 83
Table 3.7: Properties of 5HS carbon fibre fabric [151] ............................................. 84
Table 3.8: Conventional DSC experiments’ thermal protocol for non-annealed and
annealed TPU-70HS, TPU70/GNPs-NCs, TPU70/CF laminate and TPU70/CF-
MSCs. ....................................................................................................................... 103
Table 4.1: Atomic concentration of GNPs (M5, M15 and M25) as measured by
CHNS elemental analysis......................................................................................... 138
Table 4.2: Atomic concentration of GNPs (M5, M15 and M25) as measured by XPS
analysis ..................................................................................................................... 139
Table 4.3: Gel permeation chromatograph for neat TPU-70 HS and extracted TPU-70
HS matrix from its nanocomposites ......................................................................... 141
Table 4.4: TGA and DTG results for TPU-70 HS and TPU70-NCs incorporated with
5wt.% of GNPM5, GNPM15 and GNPM25 with average particle diameter 5, 15 and
25m respectively ................................................................................................... 148
Table 4.5: DSC data of first heating cycle for TPU-70 HS and TPU70/GNPs-NCs
before and after annealing ........................................................................................ 152
Table 4.6: DSC data of cooling cycle for TPU-70 HS and TPU70/GNPs-NCs before
and after annealing ................................................................................................... 152
Table 4.7: DSC second heating results for TPU-70 HS and TPU70/GNPs-NCs before
and after annealing ................................................................................................... 152
Table 4.8: Tan  for neat TPU-70 HS and TPU70/GNPs-NCs based on different
GNP sizes (M5, M15 and M25) before and after annealing .................................... 158
Table 4.9: Electrical conductivity of TPU-70 HS and TPU70/GNPs-NCs ............. 176
Table 5.1: Average properties of TPU70/CF and TPU70/GNPs/CF multiscale
composites ................................................................................................................ 182

17
Table 5.2: Storage modulus value (E’) of TPU70/CF and TPU70-MSCs at -100, 25
and 100°C ................................................................................................................. 184
Table 5.3: Average glass transition temperature of soft and hard phases before and
after annealing of TPU70/CF laminate and multiscale composites ......................... 187
Table 5.4: Comparison between actual and normalised values of tensile properties of
TPU70/CF and MSCs according to the high Vf percentage (Vf=52%) before
annealing .................................................................................................................. 197
Table 5.5: Comparison between actual and normalised values of tensile properties of
TPU70/CF and MSCs according to the high Vf percentage (Vf=52%) after annealing
.................................................................................................................................. 197
Table 5.6: Comparison between the FM and FS of current results of TPU70/CF and
MSCs before annealing with other composites incorporating various nanofillers .. 200
Table 5.7: Average in-plane and out-of-plane electrical conductivity of TPU70/CF
and MSCs ................................................................................................................. 208
Table 5.8: Comparison of in-plane and out-of-plane electrical conductivity between
the current work and previous work on composites modified by various nanofillers
.................................................................................................................................. 210
Table 5.9: Values of thermal diffusivity, thermal conductivity and heat capacity of
TPU70/CF and MSCs .............................................................................................. 213
Table 6.1: Average properties of TPU70/CF and multiscale composites ................ 225
Table 6.2: Results of Mode-I fracture-toughness test of TPU70/CF and MSCs using
MBT, including the three values of GIC for crack initiation (GIC-NL, GIC-VIS and GIC-
Max) and average GIC-Prop during crack propagation.................................................. 229
Table 6.3: Average values of GIC-Max and GIC-Prop of different epoxy/CF multiscale
composites ................................................................................................................ 236
Table 6.4: Results of low-velocity impact test of TPU70/CF laminates and MSCs at
three different impact energy levels (5 J, 15 J and 25 J) .......................................... 243
Table 6.5: Compressive strength and residual compressive strength of TPU70/CF and
MSCs before and after impact testing at 0 J, 5 J, 15 J and 25 J. .............................. 251
Table A.1: Tg and change in heat capacity (Cp) of hard phase of TPU-70 HS and
TPU70-NCs incorporating with 5 wt.% GNPs M5, GNPsM15 and GNPM25
recorded during first heating DSC scan before annealing ....................................... 284

18
Table A.2: Tg and change in heat capacity (Cp) of hard phase of TPU-70 HS and
TPU70-NCs incorporating with 5 wt.% GNPs M5, GNPsM15 and GNPM25
recorded during first heating DSC scan after annealing .......................................... 284
Table A.3: The data of Storage modulus of E’ of neat TPU-70 HS and TPU70-NCs
incorporating with 5 wt.% of various GNPs sizes (5, 15 and 25 m denoted by
GNPM5, GNPM15 and GNPM25 respectively) obtained from DMTA test before and
after annealing at three different temperature .......................................................... 284
Table A.4: Tabulated experimental data of tensile test for TPU-70 HS and
TPU70/GNPs-NCs based on the various size of GNPs(5,15 and 25 m denoted by
M5,M15 and M25 respectively) shows the values of elastic modulus, E, yeild stress,
y, tensile strength, u, and elongation at break %,u ............................................ 285
Table A.5: Tabulated experimental data of flexural test for TPU-70 HS and
TPU70/GNPs-NCs based on the various size of GNPs(5,15 and 25 m denoted by
M5,M15 and M25 respectively) shows the values of flexural strength, f, flexural
modulus Ef, and flexural strain %, f........................................................................ 286
Table B.1: Actual and normalised average values of flexural strength (FS) and
flexural modulus (FM) of TPU70/CF composites and MSCs based on the different
average diameter sizes of GNPs (5, 15 and 25 m) before annealing. The normalised
values have been calculated according to the maximum Vf = 52% ....................... 287
Table B.2: Actual and normalised average values of flexural strength (FS) and
flexural modulus (FM) of TPU70/CF composites and MSCs based on the different
average diameter sizes of GNPs (5, 15 and 25 m) after annealing. The normalised
values have been calculated according to the maximum Vf = 52% ....................... 287
Table B. 3: Average and normalised values of interlaminar shear strength (ILSS) of
TPU70/CF laminates and MSCs based on different sizes of GNPs (5, 15 and 25 m)
before and after annealing. The normalised values have been calculated according to
the maximum Vf = 47%. .......................................................................................... 287

19
List of Abbreviations

5HS Five-harness satin fabric


A-GNPs APTS modified Graphite Nanoplatelets
BDO Butanediol
CAI Compression after impact
CE Chain extender
CF Carbon fibre
CNTs Carbon nanotubes
DMTA Dynamic mechanical thermal analysis
DSC Differential scanning calorimetry
EG Expanded graphite
FAW Fibre areal weight (g/m2)
FFS Flexural failure strain
FM Flexural modulus
FS Flexural strength
G Graphite or Graphene
GF Glass fibre
GICs Graphite-intercalated compounds
GN Graphite nanosheets
GNPM15 Graphene nanoplatelets with 15 m average particle
diameter
GNPM25 Graphene nanoplatelets with 25 m average particle
diameter
GNPM5 Graphene nanoplatelets with 5 m average particle
diameter
GNPs Graphene nanoplatelets
GO Graphene oxide
HDI Hexamethylene diisocyanates
HP Hard phase
HS Hard segment
H-T Halpin-Tsai model
iGO Isocyanate graphene oxide

20
ILFT Interlaminar fracture toughness
MBT The modified beam theory
MDI Diphenylmethane diisocyanate
MH-T Modified Halpin-Tsai model
MSCs Multiscale composites
MW Molecular weight
MWCNTs Multi-wall carbon nanotubes
NCs Nanocomposites
NDT Non-destructive testing
OM Optical microscopic
PA6 Polyamide 6
PDI Polydispersity index
PEEK Polyetheretherketone
PEES polyetherethersulfone
PEI Polyetherimide
Ph-iGO Phynle isocyanate graphene oxide
PMMA Polymethylmethacrylate
PP polypropylene
PPEK Poly(phthalazinone ether ketone)
PPS Polyphenylene Sulfide
RGO Reduced Graphene oxide
RTM Resin transfer moulding
SAXS Small angle X-ray scattering
SCCNFs Stacked-cup carbon nanofibers
SEM Scanning electron microscopic
SP Soft phase
SS Soft segment
SWCNTs Small wall carbon nanotubes
TDI Toluene diisocyanate
TEM Transmission electron microscopy
TGA Thermogravimetric analysis
TPE Thermoplastic elastomers
TPs Thermoplastic polymers

21
TPU Thermoplastic polyurethane
TPU70/0.5 wt.% GNP/CF TPU70/CF laminates incorporated with 0.5 wt.%
GNPM25
TPU70/2.5 wt.% GNP/CF TPU70/CF laminates incorporated with 2.5 wt.%
GNPM25
TPU70/5 wt.% GNP/CF TPU70/CF laminates incorporated with 5 wt.%
GNPM25
TPU70/CF TPU-70 HS reinforced by carbon fibre
TPU70/CF-MSCs Multiscale composites based on the TPU-70 HS
nanocomposites and reinforced by CF
TPU70/GNPs-NCs TPU-70 HS nanocomposites incorporated with GNPs
TPU70-NCs Nnanocomposites based on the TPU-70 wt.% HS and
GNPs
TRG Thermally reduced graphene oxide
TRM Three-roll mill
TSs Thermosetting polymers
VARTM Vacuum-assisted resin transfer moulding
WAXD Wide-angle x-ray diffractions

22
List of Symbols

a Crack length (mm)


a0 Pre-crack length (mm)
TgSS The glass transition temperature of soft segments
TC Crystalline temperature
TCHS Crystallisation temperature of hard segments
Tg Glass transition temperature
TgHS The glass transition temperature of hard segments
TgMP Glass transition of a mixed-phase
Tm Melting temperature
TMMT Microphase mixing temperature
TMST Microphase separation temperature
Tdmax1 Maximum degradation temperature of the hard segment
Tdmax2 Maximum degradation temperature of the soft segment
Tdonset Initiation temperature or onset temperature of degradation
XC1 Degree of crystallinity after the first heating
XC2 Degree of crystallinity after the second heating
Tm1 first heating-melting temperature
Tm2 second heating-melting temperature
Hm1 First heating-enthalpy
Hm2 second heating-enthalpy
FSBS Interlaminar shear strength (MPa)
T Thickness of GNPs flake
EL Elastic modulus in the longitudinal direction
ET Elastic modulus in the transverse direction
GIC Mode-I critical strain energy release rate (kJ/m2)
GIC-NL GIC at the point when the load deviates from linearity in the load-
displacement curves (kJ/m2)
GIC-VIS GIC The point when the delamination can be observed on either
edge (kJ/m2)
GIC-Max GIC at the point of maximum load or at the point when the
compliance has increased by 5% (kJ/m2)

23
P Applied load (N)
Pm Maximum load
Pi Incipient or initial damage load
Et Total impact energy (J)
Er Rebound energy (J)
Ea Absorbed energy (J)
 Shape factor
 Displacement (mm)
|∆| The correction factor of crack tip rotation
vi the velocity of the impactor in m/s
vr the velocity of the rebounding impactor in m/s
ta Annealing time
Ta Annealing Temperature

24
Thesis Abstract

This project focuses on manufacturing novel carbon fibre (CF) multiscale composites
(MSCs) based on nanocomposite matrices of thermoplastic polyurethane (TPU) and
graphene nanoplatelets (GNPs). The main aim of this study is to obtain innovative
MSCs with optimised electrical, thermal and mechanical properties that make them
applicable to aerospace, transportation and electrical devices. The TPU with 70 wt.
% hard segments (HS) was filled with three different sizes of GNPs, with average
particle diameters of 5 m (GNPM5), 15 m (GNPM15) and 25 m (GNPM25), and
reinforced with CF fabric to produce MSCs. The effect of GNP size and annealing
treatment on the thermal dynamic, mechanical and electrical properties of TPU-70
HS and TPU70/CF laminate were investigated. Following these tests, the best MSCs
that could be obtained were used to study the effect of GNP content on the Mode-I
interlaminar fracture toughness (ILFT), impact-damage resistance and damage
tolerance of TPU70/CF laminates.
It was found that the tensile modulus (E) and flexural modulus (FM) of polymers-
NCs matrices were affected by both the size of GNPs and annealing treatments.
While the addition of GNPM25 yielded an impressive improvement in the E and FM
of TPU-70 HS, the annealing treatment caused a slight reduction owing to the
disruption of phase separation and restacking of the GNP flakes. However, the neat
TPU-70 HS exhibited significant improvements in the E and FM after annealing due
to the formation of microcrystalline hard domain (HD) phase separation, which
caused an increase in stiffness. The in-plane and out-of-plane electrical conductivity
and the thermal conductivity of NCs combined with GNPM15 and GNPM25 showed
a significant improvement compared to that of neat TPU-70 HS. The same results of
E and FM were reported for MSCs based on GNPM25, where the  and FM
enhanced by 19% and 105% compared with that of TPU70/CF laminates and
reduced by 11% and 4% respectively after annealing. The out-of-plane electrical
conductivity of TPU70/CF composites showed a noticeable improvement upon
incorporating GNPM15 and GNPM25. Conversely, the thermal conductivity of the
TPU70/CF laminate exhibited a significant improvement upon the addition of
GNPM5 compared to that of GNPM15 and GNPM25. This result might be attributed
to the non-uniform dispersion of GNPM5 due to its small size, leading to an increase
in local thermal diffusivity. Since the best tensile, flexural and electrical properties
were ascribed to the NCs and MSCs combined with GNPM25, these materials were
selected and different weight percentages of GNPs (0.5, 2.5 and 5) were used. The
MSCs combined with 0.5 wt.% GNPs showed a slight improvement in the G IC-Max
and GIC-Prop compared with that of the TPU-70 CF, while a high loading caused a
significant reduction in both GIC initiation and propagation. In addition, both the
impact-damage resistance and residual compressive strength of MSCs at high GNP
loading and impact energy levels experienced a significant reduction compared to
TPU70/CF laminate. To conclude, adding GNPM25 to the TPU70/CF laminate
produced MSCs with enhanced in-plane and out-of-plane electrical conductivity and
thermal conductivity. However, some mechanical properties were sacrificed, such as
ILFT, impact-damage resistance and residual compressive strength. This was
especially true at high loading, which was used to reach the percolation threshold for
GNPs with high aspect ratio.

25
Declaration

No portion of the work referred to in the thesis has been submitted in support of an
application for another degree or qualification of this or any other university or other
institute of learning.

26
Copyright Statement

I. The author of this thesis (including any appendices and/or schedules to this thesis)
owns certain copyright or related rights in it (the “Copyright”) and he has given The
University of Manchester certain rights to use such Copyright, including for
administrative purposes.

II. Copies of this thesis, either in full or in extracts and whether in hard or electronic
copy, may be made only in accordance with the Copyright, Designs and Patents Act
1988 (as amended) and regulations issued under it or, where appropriate, in
accordance with licensing agreements which the University has from time to time.
This page must form part of any such copies made.

III. The ownership of certain Copyright, patents, designs, trademarks and other
intellectual property (the “Intellectual Property”) and any reproductions of copyright
works in the thesis, for example graphs and tables (“Reproductions”), which may be
described in this thesis, may not be owned by the author and may be owned by third
parties. Such Intellectual Property and Reproductions cannot and must not be made
available for use without the prior written permission of the owner(s) of the relevant
Intellectual Property and/or Reproductions.

IV. Further information on the conditions under which disclosure, publication and
commercialisation of this thesis, the Copyright and any Intellectual Property
University IP Policy (see http://documents.manchester.ac.uk/display.aspx?
DocID=24420), in any relevant Thesis restriction declarations deposited in the
University Library, The University Library’s regulations (see
http://www.library.manchester.ac.uk/about/ regulations/) and in The University’s
policy on Presentation of Theses.

27
Dedication

This thesis is dedicated:

To the spirits of my parents, who inspired and encouraged me to be what


I am now.

To my helpful and patient husband, Julan, who devoted all his time
during my PhD studies to supporting, assisting and encouraging me.

To my beloved children and guardian angels (Fatimah, Ali and Asiah),


the lights of my life, for their understanding and help, and for providing
a suitable environment for me during my studies.

To my husband’s family, my aunt Samyiah, my brothers and sisters, who


continued to support me through my difficult times during the PhD.

28
Acknowledgment

I wish to express my gratitude to my supervisor Professor Alberto Saiani for his


dedicated support and guidance. I am grateful to Dr Arthur Wilkinson and my co-
supervisor Dr Matthieu Gresil for their suggestions and help throughout my research
project.

I wish to express my indebtedness to my sponsor, the Iraqi Ministry of Higher


Education and Scientific Research, without whose financial support I would not have
had the opportunity to complete this study. I would also like to express my
appreciation and gratitude to the Iraqi Cultural Attaché for assisting with financial
and administrative matters for the duration of my scholarship. I am so grateful to my
guarantors, Professor Haydar Al-Ethari, Professor Kadhim Alsultani and Professor
Kaisar Muslim Abd Ali at the University of Babylon, and Mr Noor Ameer Alethary,
for putting their trust in me and supporting me to receive this scholarship. I offer my
respect and thanks to the staff in the administration team, the Chemistry Department
in the College of Engineering at the University of Babylon, and to my lawyer and
father-in-law, Mr Ameer Abdullah, for their help during my studies.

My appreciation also goes to Mrs Polly Greensmith, Mr Andy Zadoroshnyj, Miss


Jasmine Fernley and Mr Ben Spencer. I highly appreciate the help and advice that I
have received during my experimental work from the staff at the Northwest
Composites Centre (NWCC), in particular from Mr Christopher Cowan, Dr Alan
Nesbitt and Ms Tracey Winspear. My special thanks go to Mr Martin Jennings and
Mr Keith Nixon in School of Chemistry at University of Manchester for their
kindness and help to finish this work.

I would also like to thank all of my colleagues and friends at the Northwest
Composites Centre and the Maths and Social Sciences Building (MSS), especially Dr
Nam Nguyen and Dr Hussien Kumer for their support and encouragement during my
PhD studies.

Last but not least, I offer my special thanks to my friends in the polymer and peptide
research group, particularly Dr Zoalfokkar Al-Obad and Dr Muayad Albozahid for
their help and support during my experimental work in my first year of study.

29
Chapter 1: Introduction

1.1 Background

The term composite materials refer to materials that consist of two phases or more
and which have a combination of properties that differs from the individual
properties of its constituents. Typically, composites are formed from a continuous
phase denoted by a matrix and a discontinuous phase denoted by a reinforcement.
The matrix is responsible for supporting and protecting the reinforcements from
chemical attack, mechanical damage and transfer of the load from one point to
another, while the reinforcement carries the load and supplies strength and thermal
stability to the composites, besides other functional properties such as electrical and
thermal conductivity [1, 2]. The material of a matrix can be ceramic, polymer, or
metal, whereas the main types of reinforcement are continuous fibre, discontinuous
fibre (short fibre), whiskers (elongated single crystal), and particles [2, 3].
Conventional fibre-reinforced polymeric composite materials are the most common
among the other reinforcements, owing to their extraordinary mechanical properties
and high strength-to-weight ratio. These superior properties have led them to replace
aluminium and steel in many applications, including the military, automobile,
aerospace, marine and energy sectors, in technology such as wind turbine blades.

Carbon fibres are extensively utilised in the manufacture of polymer-matrix


composites due to their good mechanical, electrical and thermal properties, their high
specific strength and stiffness, and high fatigue strength, in comparison with other
reinforcing fibre materials (aramid, glass, or natural fibre). As such, they can be
considered to have a great ability to improve polymer properties [4]. However, the
effective ability to use CF composites in this application is limited due to its
susceptibility to accidental low-energy impact, compressive loads and interlaminar
fracture [5-9]. This susceptibility arises from its weakness in out-of-plane properties,
particularly low damage tolerance, poor interlaminar fracture toughness and impact-
damage resistance. These weaknesses are in turn owing to its combination with
brittle polymer matrices such as epoxy, which is normally used in CF laminates [1, 5,
10]. The most significant damage that forms in CF laminates, which overshadows all
other weaknesses in brittle matrices besides weak fibre-matrix adhesion, is

30
delamination. Delamination is considered the most dangerous type of damage that
forms within CF composites, as it can lead to a great reduction in the strength and
provoke a catastrophic failure during composites’ structural service. To control this
problem, different methods have been employed to improve the through-thickness
properties of CF composites via exploiting reinforcing elements, such as 3D textiles
or stitching [11]. Yet, these methods neither improve the intrinsic interface of CF-
composite structures nor the properties of matrices [9]. Therefore, a substantial
amount of research has recently taken an interest in using thermoplastic polymer as a
matrix with CF or glass fibre reinforcement due to their advantages such as high
toughness, impact resistance and short processing time in comparison with the
thermosetting polymer. As such, composites based on thermoplastic matrices can
overcome the weakness in composite laminates that arises from using brittle matrices
[12-14]. Like previous studies on thermoplastic composites [12-14], this study is
focused on the use of thermoplastic polyurethane (TPU) as a matrix with CF owing
to its individual properties. TPUs possess a unique segmented structure which gives
them excellent resilience and toughness, as well as good abrasion resistance.
Furthermore, TPUs have adhesive properties which help them to increase the
adhesion between CF and a matrix without any requirement for additional
components [15].

On the other hand, TPUs show low stiffness and weak heat resistance in addition to
their low thermal and electrical properties combining them with short CF were
proved as an effective method to enhance their mechanical, thermal and electrical
properties [16-18]. As such, using TPUs with fabric CF can consider as innovative
composites at the time of this study since must researches were dealt with TPUs
combined with short CF [16-18]. Otherwise combine TPUs (low thermal and
electrical properties) with CF (high thermal and electrical properties) can affect the
overall electrical and thermal properties of the resultant composites in both the in-
plane and, to a much greater degree, in the out-of-plane direction. This is because of
the role of the matrix in controlling the properties of CF composites in the out-of-
plane direction, while the CF dominates the in-plane properties. Hence, to improve
the electrical and thermal properties of a matrix and then to enhance the out-of-plane
properties of composites, conductive nanofillers such as carbon nanotubes (CNTs)
and graphene or graphene nanoplatelets (GNPs) can be incorporated in polymer

31
matrices. In recent years, different studies have attempted to improve the through-
thickness properties of fibre/composites based on thermoplastic matrices via the
incorporation of nanofiller particles [9, 19, 20].

GNPs are desirable to work with due to their straightforward and low-cost
production in comparison with single-layer graphene and carbon nanotubes (CNTs)
[21, 22]. Commercial GNPs are available in various grades, especially according to
their diameter, which can vary from nanometres to hundredths of micrometres [22].
Advancements in nanotechnology that is based on modifying polymer matrices using
GNPs and the optimised electrical, thermal and mechanical properties that were
gained in the resultant nanocomposites [23-29] are of interest to the current study of
the modification of TPUs using GNPs. The ability to obtain GNPs of different sizes
enables this study to focus on studying the effect of GNP sizes, rather than their
concentrations, on the properties of TPU nanocomposites (NCs). The proposed
concept in this study is to form NCs based on TPUs as a matrix and reinforced with
different sizes of GNPs. The filled and unfilled TPU will then be used to fabricate
CF composites and innovative multiscale composites (MSCs). As a result, by
conducting a comprehensive study beginning with the synthesis of TPU-70 HS and
TPU70-NCs, followed by the manufacture of NCs and MSCs, through to testing and
discussion, the results will contribute an understanding of the effect of GNP sizes on
the properties of TPU70-NCs and MSCs.

1.2 Global Market of CF composites

Recently, the demands of CF composites are significantly increased due to the


interfere of CF composites in different applications such as automotive, wind
turbines, sports, aerospace and marine [30]. In United State, the market size of
carbon fibre composites was worth over 16 billion dollars in 2014 and expected to
increase the demands of carbon fibre composites to be 290 k tons in 2024 [31]. This
growth in demanding of carbon fibre was due to the increase in the adaption sector of
aerospace and automotive which have been driven the industry until 2024 [31].

On the other hand, the usage of CF composites in the automotive sector in 2016
captured a dramatic share over a percentage of 20 and higher than 25 k tons in
demand [30]. This percentage is expected to grow by 13 % in 2024 due to the low
price of polymer matrices of CF composite and the superior properties of resultant

32
CF composites such as superior strength along with light weight [30, 31]. Figure 1.1
illustrates the comparison of the usage of carbon fibre composites in 2016 and their
expected one in 2024 in United State region.

Figure 1.1: The market of carbon fibre composites by the end of 2016 and 2024 in
United States estimated in millions dollars [31]
For the matrices used to fabricate CF composite, the polymer matrices recorded the
highest usage in 2013 arounds 64% compared with others matrices such as ceramics
and metals according to report presented by Kraus and Kühnel in United States
regions [32] as illustrated in Figure 1.2.

Figure 1.2: The revenue of matrices of carbon fibre composites in 2013 in United
States regions[32]

33
From this figure, the usage of thermosets polymers as a matrices recorded a percent
around 76% from 64% of all polymers types in 2013, while the thermoplastic
matrices including TPU recorded around 24% from the total percent [30, 32]. The
results of this report confirm the suggestion about using thermoplastic matrices to
combine with CF is a new research in order to overcome the weakness in thermosets
matrices which was mentioned in previous section.

1.3 Research Aims and Objectives

The main aim of this study is to fabricate innovative MSCs with optimised electrical,
thermal and mechanical properties that prepare them for use in different applications,
such as in the electromagnetic shielding, aerospace, aircraft, marine and
transportation industries. The MSCs are fabricated using TPUs as a matrix with the
incorporation of different sizes of GNPs, and then the NCs reinforced with CF fabric.
After that, the optimum GNP sizes that form NCs and MSCs with superior properties
will be selected and used to study the effect of GNP concentration on the
interlaminar fracture toughness, impact resistance and damage tolerance of MSCs.
This procedure will be followed in order to understand if the improvement in
through-thickness electrical properties will lead to sacrificing the most important
mechanical properties of CF composites, which are delamination resistance and
damage tolerance.

The objectives of this research can be summarised as follows:

1. Synthesise TPU with 70 wt.% HS using a pre-polymer technique.

2. Synthesise TPU70-NCs via an in-situ polymerisation process.

3. Ascertain the best casting procedure for both TPU-70 HS and TPU70-NCs.

4. Fabricate TPU70/CF laminates and TPU70-MSCs via finding the best


processing conditions for the compression-moulding process.

5. Study the effect of annealing treatment and GNP sizes on the thermal,
dynamic and mechanical properties of neat TPU-70 HS and TPU70-NCs.

6. Study the effect of GNP sizes on the thermal and electrical properties of neat
TPU-70 HS and TPU70-NCs.

34
7. Study the effect of annealing treatment and GNP sizes on the dynamic and
mechanical properties of TPU70/CF laminates and TPU70-MSCs.

8. Investigate the effect of GNP sizes on the thermal and electrical conductivity
of TPU70/CF laminates and TPU70-MSCs.

9. Find the best GNP sizes that match with TPU-70 HS matrices and TPU70/CF
composites, leading to better improvement in the electrical and mechanical
properties of resultant NCs and MSCs.

10. Investigate the effect of the content of the nominated GNP size on the
delamination resistance, damage resistance and damage tolerance of TPU70-
MSCs.

1.4 Thesis Structure

This thesis consists of seven chapters. Chapter 1 has served to summarise a brief
introduction of the background, the aims and objectives of this research and the
structure of this thesis.

Chapter 2 provides a review of the literature on TPUs and the factors that affect their
properties. Brief descriptions of graphene and GNPs are presented in this chapter, as
well as the methods that have been used to disperse nanoparticles within polymer
matrices. In addition, this chapter indicates the best dispersion method that has been
used previously to disperse GNPs within a TPU-70 HS matrix and which will be
adopted in this research. The effect of graphene and its derivatives on the
mechanical, electrical and thermal properties of polymer matrices is discussed in
detail in this chapter. The fabrication of MSCs and the effect of nanoparticles on
their mechanical, electrical and thermal properties are discussed in detail by referring
to various previous studies of different MSC systems.

Chapter 3 describes the methodology that is applied in this study, including a


description of the raw materials of synthesised TPUs, TPU-NCs and MSCs. The
experimental techniques that are used to investigated NCs and MSCs are also
described, as well as the intrinsic principles of each technique.

35
Chapter 4 focuses on the characterisations that have been carried out of both
graphene nanoplatelets and TPU70-NCs, including a discussion of the results of the
effect of GNP sizes and annealing treatment on the thermal, dynamic and mechanical
properties of NCs. A discussion of the results of the effect of GNP sizes on the
thermal and electrical conductivity of TPU70-NCs is also given in this chapter.

Chapter 5 discusses the results of the effect of GNP sizes and annealing treatments
on the dynamic and mechanical properties of TPU70/CF laminates and TPU70-
MSCs. The results of the effect of GNP sizes on the electrical and thermal
conductivity of TPU70/CF and MSCs are also discussed in this chapter.

Chapter 6 discusses the results of the effect of weight percentages of nominated GNP
sizes on the interlaminar fracture toughness, impact-damage resistance and damage
tolerance of TPU70/CF laminates and MSCs.

Chapter 7 includes the conclusion of all of the results of the project and suggestions
for future work.

36
Chapter 2: Literature Review

2.1 Thermoplastic and Thermoplastic Elastomers

Over the last forty years, thermoplastic polymers (TPs) have attracted substantial
attention for their use as a matrix for composite materials due to their superior
properties to thermosetting polymers. These main important properties are excellent
impact resistance; high damage tolerance, which results from good toughness,
recyclability and weldability [33-35]. The main disadvantage of thermoplastic
polymers in comparison with thermosetting is their low stiffness and then low
through thickness properties. Thermoplastic polymers can be amorphous or semi-
crystalline, depending on both the polymer structure and the processing techniques
used [36]. Examples of amorphous TPs are polyetherimide (PEI), polyvinyl chloride
(PVC) and polystyrene. These types of thermoplastic polymers are transparent, have
low density, low shrinkage, a broad softening point and low chemical resistance. In
contrast, semi-crystalline TPPs such as polyphenylene sulfide (PPS), polypropylene
(PP), polyetheretherketone (PEEK) and polyamide (PA) are opaque and possess
high shrinkage, high density and high chemical resistance good fatigue and wear
properties [37].

Thermoplastic elastomers (TPEs) are part of another polymer family that has two
phases: the first one is amorphous and soft at ambient temperature, while the other is
hard and glassy and/or crystalline and can be bonded together chemically by block or
graft polymerisation. In general, the soft phase (SP) gives the TPEs flexibility and
elasticity, whereas the hard phase (HP) provides strength to the system and
represents physical cross-linking. Hence, these materials behave like rubber at
ambient temperature but, when heated above the hard-phase melting temperature or
when the hard phase dissolves in a solvent, they can flow and form by the same
processes as thermoplastic (for example, extrusion, injection moulding, compression
moulding, blow moulding, etc.). They can also be recycled [38, 39]. Compared with
conventional elastomers, TPEs have many advantages; the main one is their low
energy cost during processing and the availability of standard uniform grades [36]. In
general, both the percentages of hard thermoplastic and soft rubbery segments have
an effect on the mechanical properties of the resulting TPEs. Namely, hard segments

37
directly affect the tear and tensile strength, fluids and chemical resistance, and the
processing temperature, while soft segments affect the hardness, flexibility, the limit
of the service at low temperature, and the tensile or compression set [36].

Thermoplastic polyurethanes (TPUs) are one of the commercial classes of TPEs.


These are versatile materials which have recently begun to play an important role in
numerous industrial sectors, ranging from the footwear industry to biomedical
material applications [40-42]. This is because of their unique properties, which
include good mechanical properties, vibration damping, their ability to resist
chemicals, grease, oil, and abrasion [38, 43, 44]. Furthermore, another important
factor relates to the possibility of tailoring the structure of TPU by varying the type
and ratio of the raw components during the synthesising process [45]. Therefore,
TPU has attracted many researchers to use it to produce composite or nanocomposite
materials [46-54]. Besides the advantages of TPU indicated earlier, research interest
in this material has arisen due to its simple synthesis protocol, low melting
temperature (175°C) and high processing speed [46, 47].

2.1.1 Thermoplastic Polyurethane

Thermoplastic polyurethanes (TPUs) are typically based upon linear segmented


copolymers which consist of alternating statistical distribution of soft segments (SS)
and hard segments (HS), as shown in Figure 2.1 [45].

(A)

(B)

Figure 2.1: Schematic representation of TPU structure: (A) repeat hard and soft
segment units, and (B) alternating structure of hard segments (HS) and soft segments
(SS) [45]

38
The SS are flexible segments with poor polarity that have a low glass-transition
temperature and high molecular weight, coming from a polyol (alcohol
functionalised chain) which is typically either a saturated aliphatic polyester or
polyether. On the contrary, HS are polar, more rigid segments and usually have a
glass transition above room temperature, thus they are known as HS as a result of
their stiffness in normal ambient conditions [41, 52, 55]. HS are formed by the
reaction between a difunctional (sometimes multifunctional) isocyanate such as
diphenylemethane-4, 4-diisocyanate (MDI) and a chain extender (CE) of low
molecular weight, mostly diamines or diols.

Generally, the HS of a TPU structure can be either aromatic or aliphatic depending


on the type of isocyanates. Aliphatic isocyanates, such as hexamethylene
diisocyanates (HDI), are less reactive than aromatic diisocyanate. It has been
reported that aliphatic diisocyanate is suitable for use in biomedical applications due
to its ability to degrade into non-toxic decomposition products [56, 57]. On the
contrary, aromatic diisocyanates such as diphenylmethane diisocyanate (MDI) and
toluene diisocyanate (TDI) are suitable for use in many different applications that
require flexibility, toughness and strength [45]. MDI is the most common aromatic
diisocyanate that has been used to synthesise TPUs owing to its low cost, high
reactivity, ability to form microphase-separated morphologies, ordered hard domain
and useful mechanical properties resulting from its symmetrical structure [44, 56,
58].

On the other hand, a TPU structure can be based on polyester or polyether polyol, as
mentioned earlier. The TPUs based on the polyester polyol possess high tear and
abrasion resistance and exceptional resistance to non-polar fluids, while those based
on polyether polyol have excellent thermal and hydrolytic stability, excellent
resilience and high resistance to oxidative attack [42]. As pointed out previously, CE
can also be formed of diamines or diols; hence urea HS results from using diamines,
whilst urethane HS forms as a consequence of using diols in the CE [43, 45].
Furthermore, the TPUs that result from using a CE of diamines are only suitable for
solvent casting, whereas the ones produced by using a diol CE usually have thermal
processing abilities [59].

In general, the macromonomer chain of polyols is oligomeric and terminated by


hydroxyl groups (-OH), while the small molecule of CE is either terminated with

39
hydroxyl or amine end groups. The easy reaction between diisocyanate groups and
hydroxyl groups of polyols on one side and CE on the others leads to the production
of segmented polyurethane. This segmented polyurethane can be linear if
bifunctional polyol, isocyanate and CE are used, and branched if polyol, isocyanate
and occasionally CE are multifunctional [60].

The process of synthesising TPUs can be carried out in the laboratory or in small-
batch production using two methods, consisting of one step (one-shot) or two steps
(pre-polymer). In the one-shot method, all the reactants (polyol, diisocyanate and
chain extender) are mixed together and allowed to react in one step. The TPUs
produced from this method contain HS with random lengths due to the ability of the
CE to react with diisocyanate before the polyol [38, 61, 62]. In contrast, a pre-
polymer method is accomplished in two steps; in the first one, the polyol reacts with
a large addition of diisocyanate to produce a pre-polymer. In the second step, the pre-
polymer reacts with the chain extender and extra diisocyanate to form high-
molecular-weight segmented TPUs [38, 59, 62]. The TPU structure that results from
the pre-polymer method is more regular than the one produced in the one-step
method. This is due to the role of pre-polymer method in capping polyol and
diisocyanate together and connecting them with CE, leading to regular HS and SS
distribution. As a result of the regularity in the HS-SS sequence, the TPU materials
may have superior mechanical properties, because it is much easier to create physical
cross-linking points in aggregated or crystalline HS [45, 59].

Generally, the properties of TPUs are attributed to the formation of urethane linkage
(Figure 2.2) in the molecular backbone. The urethane linkage represents the principal
chemical reaction in the TPU synthesis process, which occurs due to the reaction of
isocyanate with hydroxyl groups [44, 60, 63].

Figure 2.2: Urethane linkage [60]


In polyurethane, both physical and chemical cross-linking are found. For example,
physical cross-links can be attained through hydrogen bonding and the formation of a
hard domain, achieved by the phase-separated system of the soft and hard segments.

40
These physical cross-links can be broken by using heat or a solvent and reformed by
cooling or evaporation [39, 43]. In contrast, chemical cross-linking is obtained by tri-
or multifunctional constituents and is not easy to destroy by heating or dissolution
except in some special cases of labile chemical groups, leading to the production of
an irreversible network [43, 45]. However, the SS connects to the chemical cross-
links and is responsible for the flexibility of TPU. Owing to the low glass transition
temperature (Tg) and melting temperature (Tm) of SS compared with HS, upon
heating to temperatures in excess of the melting point of HS, the TPU polymer
becomes a homogeneous viscous mixture. As a result, TPU polymer can be formed
using normal melt-processing methods for thermoplastic materials such as extrusion,
injection moulding, compression moulding, blow moulding, etc., as has been
indicated previously [38, 43, 62]. On the other hand, below the Tg of SS, their brittle
nature overcomes the TPUs’ polymer behaviour and its useful elastomeric
characteristics will be lost [42].

In a TPU structure, hydrogen bonding has a significant effect on their properties


owing to bonds forming between NH groups in urethane linkages that act as a donor
and the acceptor of hydrogen bonds, which is either the carbonyl groups (C=O) of
the urethane group (hard segments) or the soft segments (carbonyl in ester or oxygen
in ether) [45, 64]. Generally, the nature of the urethane group tends to build self-
association by forming hydrogen bonds in either parallel or anti-parallel patterns.
Since the concept of hydrogen bonds has been utilised to estimate the degree of
phase separation [65], for many decades the type of H-bond in PU structures has
been the topic of several investigations. For more detail, four types of H-bonds in a
TPU structure based on the polyether polyol and a diol CE are illustrated in Figure
2.3.

Figure 2.3: Schematic representation depicting four types of H-bonds relating to


polyether-based polyurethane, where (I), (II) and (III) are urethane–urethane
interactions, while (IV) is a urethane-soft segment interaction [66, 67]

41
These four patterns represent the H-bonding that forms between an NH group and I)
C=O groups, II) urethane alkoxy oxygens, III) an NH group leading to form NH---
NH bonds and IV) an ether (C-O-C) group. The first three types represent urethane-
urethane interactions, while the fourth one deals with urethane-soft segment
interactions [66, 67].

2.1.1.1 Morphology of TPU Structure


As a consequence of the thermodynamic incompatibility between soft and hard
segments, a microphase-separated structure is formed, making it is possible to
distinguish the domain structures of HS and SS, as illustrated in Figure 2.4 [43, 45,
60, 63].

Hard segment

Soft segment

Hard domain

Soft domain

Figure 2.4: Schematic representation of the microphase morphology of TPU


structure [63]
The microphase separation and the characterisations of the domain morphology
strongly affect the properties of the final TPU materials [45, 68]. At low weight
percentages of HS, the formation of a discrete hard domain represents physical cross-
linking points and/or rigid filler that are dispersed in an amorphous SS matrix [45,
63]. When the amount of HS is increased, consequently, the interconnectivity of hard
domains also increases, leading to a transition from a separate hard domain to a
continuous one [63, 69].

TPUs are similar to other polymers in that they are strongly influenced by changes in
heat. Therefore, it becomes crucial to understand the effect of temperature variation

42
on the phase separation and properties of TPUs. Usually, the thermal transitions of
segmented PU can be detected using differential scanning calorimetry (DSC),
through the use of which five regions of thermal transitions can be observed as
illustrated in Figure 2.5.

Figure 2.5: Thermal transition of segmented PU by using DSC techniques [70]


These regions are: the glass transitions of the soft phase (TgSP), the glass transition of
the hard phase (TgHP), the glass transition of a mixed phase (TgMP), the microphase
separation temperature (TMST), and finally the temperature of microphase mixing
temperature (TMMT). The endothermic peak at high temperature denoted by TMST and
TMMT is related to the segregation of the SS and HS of the TPU structure. That is,
TMST is detected if a complete segregation of SS and HS occurs, while the
endothermic peak donated by TMMT occurs when the segregation of SS and HS is
only partial [70-73].

Koberstein and Russell [74] studied the endothermic transition of segmented PU


elastomer using a DSC scan and small-angle-X-ray scattering (SAXS). The
segmented PU that was used was based on the MDI and butanediol as hard segments
and oxyethylene end-capped polyoxypropylene as soft segments. Two distinct
endothermic transitions of segmented PU were observed: an intermediate-

43
temperature endotherm (TII) and a high-temperature endotherm (TIII). The TII was
attributed to the onset of micro-domain mixing between monocrystalline hard
segments and soft segments, and the transition temperature was associated with the
transition from the ordered phase to the disordered one, while TIII was attributed to
the melting of microcrystalline HS. The morphological changes of segmented PU
elastomers obtained during the DSC scan are shown in Figure 2.6.

Figure 2.6: Schematic representations of the morphological changes of segmented


polyurethane elastomers obtained through DSC scans (a) below the microphase
mixing temperature (TII), (b) between the melting temperature and TII, and (c) above
the temperature of the microcrystalline hard domain (TIII) [74]
In this figure, the schematics a, b and c represent respectively the morphologies of
PU at applied temperatures lower than the microphase mixing temperature (TII), at
applied temperatures between TII and the melting temperature of the microcrystalline
hard domain (TIII), and at applied temperatures higher than TIII. The researchers
concluded that both the TII and TIII increased in line with increasing content of hard
segments and soft segments.

On the other hand, the phase separation of TPU structures is highly dependent upon
the formation of hydrogen bonds between urethane linkages, synthesis conditions
and the manufacturing process [41, 45]. In reality, it is difficult to have a complete
phase separation because there is a certain amount of hard or soft segments that
remains in each domain [75-77]. Many factors affect the morphology of phase

44
separation and, in turn, the properties of the TPU, such as the amount of soft and
hard segments, the ratio between isocyanate (NCO) and hydroxyl groups, the thermal
history of the TPU, and synthesis processing methods (such as pre-polymer or one-
shot methods). In addition, the annealing treatment has a significant effect on the
mechanical properties of TPU due to its effect on the formation and degree of phase
separation. As such, the morphology of phase separation and the variation in the
properties and composition of TPU have attracted the attention of researchers for
decades owing to their economic importance. The following two sections will focus
on existing research that has dealt with the effects of HS contents and the effects of
annealing treatment on the morphology and structure of TPU.

A) Effect of weight percentage of Hard Segments

One crucial factor which affects TPU morphology and its properties is the chemical
composition and the ratio of soft and hard segments, since these segments are
responsible for the flexibility and hardness of TPU polymers. One of many research
projects that has studied these factors was performed by Lee et al. [78]. The authors
used the two-step polymerisation process to synthesise block copolymer TPU based
on HS made of 4,4-MDI and extended with 1,4-butanediol and SS made of poly
(tetramethylene glycol) (PTMG). Contents of HS ranging from 20 to 50 wt.% were
used during their work. They concluded that the hard segment content, degree of
crystallinity and phase separation have a great effect on the mechanical properties of
copolymer PU. It was reported that PU with high HS content experienced high
elastic modulus and tensile strength but lower elongation at break. They attributed
this to the fact that the hydrogen bonding in the urethane linkage of HS and between
the SS and HS increased with increasing the HS content, making the TPU hard to
stretch under the tensile test. Therefore, the samples with high HS content exhibited a
high elastic modulus but failed in their low elongation [78].

The same results were reported by Al-Obad [47], who studied the effect of HS
contents on the tensile properties of TPU before and after annealing. The researcher
used two weight percentages of HS (70 wt.% and 90 wt.% HS) based on HS made
from 4,4-MDI and 1,5 butanediol as a chain extender, and the SS was made from
polyol (polyethylene glycol-block-polypropylene glycol-block-polyethylene glycol).
It was demonstrated that an increase in the HS content of the TPU structure led to an

45
increase in the elastic modulus and tensile strength, but reduced the elongation at
break. The findings of Al-Obad’s study [47] were compatible with the results
obtained by Albozahid [46], who used the same chemical composition of HS and SS
of TPU of Al-Obad’s study [47] but different hard-segment ratios, which were 60
wt.%, 70 wt.%, and 80 wt.%. Albozahid [46] studied the effect of hard segment
content on the tensile properties of TPU. In spite of the similarities of the study with
the work conducted by Al-Obad [47], his results confirmed that the TPU-70 HS
showed intermediate improvement in tensile properties before and after annealing
compared with TPU-60 HS and TPU-80 HS. When the TPU-60 HS failed at a lower
tensile strength and high elongation at break, the TPU-80 HS ruptured at
approximately the same tensile strength of TPU-70 HS but with a very low
elongation at break, as illustrated in Figure 2.7.

(a) (b)

NON-ANN ANN NON-ANN ANN NON-ANN ANN NON-ANN ANN NON-ANN ANN NON-ANN ANN NON-ANN ANN NON-ANN ANN

(c)

NON-ANN ANN NON-ANN ANN NON-ANN ANN NON-ANN ANN

Figure 2.7: Comparison of tensile test results (modulus, yield strength and
elongation at break) of TPU based on the different hard segments: a) TPU with 60
wt.% HS; b) TPU with 70 wt.% HS; and c) TPU with 80 wt.% HS before and after
annealing [46]

46
B) Effect of Annealing Treatment on the Morphology of TPU structure

Annealing treatment has a great effect on the mechanical properties of TPUs due to
its effect on the formation of phase separation of SS and HS in segmented TPU
structures [79-81]. The annealing treatment is considered a crucial process in
industrial fields that can contribute to attaining a product with desirable properties.
However, the complexity of structures of TPUs due to its variation after annealing
makes them difficult to understand. As a result, many research groups have studied
the effect of annealing on the thermal transition of TPU structures, as exemplified by
the work of Saiani et al. [71], who studied the influence of annealing time (ta) using
DSC and wide-angle X-ray scattering (WAXS) on a different set of TPUs based on
different weight percentages of hard segments ranging from 50% to 100%. They
used MDI and a chain extender of 2-methyl-1,3-propanediol (MP-Diol) as hard
segments and polypropylene oxide end-capped with ethylene oxide as SS. Their tests
enabled them to observe the endothermic transition of high temperature, disruption in
the ordering structure of the hard phase under certain annealing conditions, and the
formation of microphase separation between the SS and HS. It was found that
annealing for different times at 120 oC gave two endothermic peaks, Tm and TMMT.
Tm represents the lower melting temperature endotherm, which is attributed to the
ordering structure presented in segregated hard domains, while the higher melting
temperature endotherm is related to the microphase mixing temperature (TMMT) for
soft and hard segments, as shown in Figure 2.8.

Figure 2.8: DSC results for PU-65%HS at 120 °C at different annealing times (ta)
obtained at a heating rate of 20 °C/min [71]
47
The researchers revealed that both Tm and melting enthalpies shifted to a higher
temperature with increasing ta, and demonstrated that the annealing treatment caused
an increase in phase separation and formation of ordering hard domains. The results
of WAXS confirmed the results of DSC test relating to increase the formation of
microphase separation between soft and hard segments with increasing annealing
time which gave further support to their study. Additionally, the results of melt-
quenched samples for HS (less than 65%) showed that a homogeneous mixed-phase
was achieved, whereas for higher percentages of HS two phases were attained (the
HS phase coincided with the mixing phase for both SS and HS) [71].

In a later study conducted by Saiani et al. [73] carried out using the same set of
samples of linear TPU used in their previous study [71], an investigation of the
morphology and structure of TPU was performed using SAXS and transmission
electron microscopy (TEM). Their results confirmed the findings of the previous
study and found that the morphology of samples with an HS content greater than
65% consisted of SS domains in a HS matrix. These results were confirmed with
their last DSC results [71] and hypothesised melt-quenched samples with a two-
phase structure: pure HS plus a mixed-phase (hard plus soft segments) with an
amount of HS of around 65 wt.%, as shown in Figure 2.9 a.

(a) (b)

(c)

Figure 2.9: Schematic representation of a morphological model illustrating the TPU


systems for different phases. (a) Melt-quenched samples, (b) microphase separated
samples, (c) microphase separated samples under mechanical stress [73]

48
After annealing, the mixed-phase of both the hard and soft segments might undergo
phase segregation resulting from the annealing process, as shown in Figure 2.9 b.
Furthermore, to give more support to this interpretation, the intensity of all SAXS
scattering exhibits a reduction with increasing hard segment content. Finally, the
researchers observed that the soft segments tended to align into a lamellar
morphology under mechanical testing [73] (see Figure 2.9 c).

Another study was published by Nallicheri and Rubner [79], who investigated the
effect of annealing treatment on the thermal transition and mechanical properties of a
series of TPU polymers. These series of TPU were based on hard segments made of
MDI or HDI and extended with either 2,4-hexadiyne-l,6-diol or 5,7-dodecadiyne-
l,12-diol and SS made of poly(tetramethylene oxide) (PTMO) with a MW of 2000
g/mol or 1000 g/mol. They synthesised TPU using a pre-polymer method. The study
revealed that the ultimate tensile strength of the materials was strongly influenced by
their MW and thermal history. It was found that the mechanical properties exhibited
a manifest improvement with an increase in MW. Furthermore, it was demonstrated
that the annealing treatment led to a significant enhancement in the mechanical
properties. They suggested via thermal analysis using DSC that the annealing
treatment caused an increase in the cohesion forces between hard segments,
expanding the integrity of the hard domain structure. They found that the
endothermic peak associated with the hard domain shifted to a higher temperature,
indicating an increase in the ordering of the hard domain. Thereby, the ultimate
tensile strength was found to improve after annealing treatment due to the increase in
the cohesion of HS and the ordering of the hard domain [79].

Another study of the effect of annealing treatment on the tensile properties of TPUs
was carried out by Yanagihara et al. [82]. They used a pre-polymer method to
synthesise TPU based on HS made of MDI and 1,4-butanediol (1,4-BD) and a SS
made of polyester glycol, with a weight percentage of 43 wt.% of HS in the TPU
structure. Firstly, they made a sample using a compression-moulding process at 195
and 120 °C. Afterwards, the moulded samples were immediately quenched in ice
water at 0 °C and placed in the oven for annealing at temperatures ranging from 25 to
145 °C for 16 hours. To achieve their purpose, they used SAXS, tensile testing and
dynamic mechanical thermal analysis (DMTA). It was concluded that the elastic
modulus of the annealed sample reduced with an increase in the annealing

49
temperature (Ta), in spite of the increase in size and fraction of the hard domain after
annealing and with increasing Ta. The researchers ascribed their results to the
formation of an HS-rich domain, as shown in Figure 2.10.

(a) Just after the melt-quench (b) Annealing MQ at 25 C

(c) Annealing MQ at higher Ta

Figure 2.10: Schematic representation of the development of an HS-rich domain in a


TPU structure at different annealing temperatures (Ta): (a) sample directly after
melt-quenching (MQ), b) MQ sample annealed at 25°C, and c) MQ samples
annealed at higher Ta [82]
In this figure, (a) represents the structure after melt-quenching (MQ) directly; which
led to forming a hard segment-rich domain as loosely packed spherical, due to
inadequate time being given after the MQ process, resulting a low elastic modulus.
On the other hand, (b) illustrates the MQ structure after annealing at 25 °C when the
HS-rich domain converted to a large ellipsoid domain with a fringed-micelle-like
structure; moreover, it became denser and more aligned owing to the H-bonding
amongst urethane groups, resulting in an increase in elastic modulus. Finally, Figure
2.10 c, indicates the structure after annealing at higher Ta. The authors explained that
the increase in Ta led to a decrease in the density and alignment of the HS chains due
to the reduction in strength of the H-bonding, causing a reduction in the elastic
modulus of the annealed samples [82].
50
Al-Obad [47] performed a study of the effect of annealing treatment on TPU based
on two different weight percentage of HS (70 and 90 wt.% as mentioned previously).
He chose 80 C to do annealing treatment at different period started from 1 day to 7
days. The author used DSC and WAXS techniques to study the effect of annealing
time on the crystallinity of TPU. The results demonstrated that the best time to get
higher crystallinity of TPU-70 HS is 3 days, while 4 days is the best time of
annealing TPU-90 HS. Therefore, this study adopted annealing TPU-70 HS and its
composites at 80 C and for 3 days as recommended by Al-Obad [47].

2.2 Nanocomposites and Nanocomposite Matrices

2.2.1 Graphitic Forms of Carbon

Carbon is a unique element that has the ability to form different types of materials
through bonding with its own atoms or with other elements, such as hydrogen,
nitrogen, oxygen and sulphur [83]. Carbon elements can be found in crystalline
forms, such as diamond and graphite. Diamond is the hardest material in the universe
and consists of carbon atoms, arranging and bonding to each other via covalent
bonds to form a tetrahedral lattice. On the contrary, graphite is a very soft material,
and consists of a series of stacked parallel graphene layers. Each layer comprises sp 2
hybridised carbon atoms covalently bonded to three other atoms, forming a planar
hexagonal ring. Each layer is bonded to the neighbouring layer via van der Waals
forces when its hybridised fourth valence electron is paired with another
neighbouring delocalised electron [83, 84], as illustrated in Figure 2.11.

Figure 2.11: Graphite structure illustrating the sp2 hybridised carbon atoms bonded
in continuous hexagons [85]

51
Until 2004, there was an ongoing debate surrounding isolating a monolayer of
graphene from graphite, with the prevailing assumption that a 2D crystal is
thermodynamically unstable and cannot exist [86]. In 2004 at Manchester University,
Geim and Novoselov [87] refuted this supposition when they succeeded in peeling
off a single graphene sheet from a graphite structure using the ‘Scotch tape
technique’. Their discovery is considered a breakthrough in scientific research and
industrial fields due to the potential application of graphene in various areas.
Graphene sheets are the basis of all graphitic forms of carbon, as they can be stacked
to form 3D graphite, wrapped to form 0D fullerenes (buckyballs) and rolled up to
produce 1D nanotubes [88], as shown in Figure 2.12.

Figure 2.12: Graphene as the mother of all graphitic forms: when stacked, it forms
3D graphite, or can be wrapped into 0D fullerenes and rolled into 1D nanotubes
[86, 88]

2.2.2 Graphene Nanoplatelets (GNPs)

Graphene nanoplatelets (GNPs) are a type of carbon nanofillers, consisting of (10-


30) stacked 2D graphene sheets, which possess outstanding electrical, thermal and
mechanical properties [22, 89]. GNPs are also known as graphite nanosheets (GN),
exfoliated or expanded graphite (EG), and graphite nanoflakes (GNF). They are
usually attained by exfoliating graphite flakes, leading to form GNPs with a
thickness ranging from less than 0.35 to 100 nm [85]. GNPs possess a high surface
area and high aspect ratio due to their diameter, which is measured on microscale in
comparison with their thickness. The diameter and thickness of GNPs can be altered

52
and controlled via various techniques, such as oxidation, intercalation, microwave
irradiation, heat treatment, and ultrasonic treatment [22, 90].

GNPs can be synthesised using chemical vapour deposition (CVD) and arc
discharging; however, the two preferred fabrication methods are mechanical milling
and graphite intercalation chemistry [22, 85, 90]. The mechanical milling method
starts to produce GNPs by breaking up graphite flakes, resulting in large GNP
particles with non-uniform size distribution. To solve this problem, the intercalation
method can be followed. In the intercalation method, large GNPs undergo the
separation of a graphene layer via exposure to acid. Generally, intercalation is carried
out by immersing natural graphite in a chemical solution composed of a concentrated
sulfuric acid (H2SO4) and various oxidisers such as HNO3, O3, and KMnO4. The
GNPs that result from intercalation are called graphite-intercalated compounds (GIC)
or expandable graphite [22]. A schematic of the modification of graphite is illustrated
in Figure 2.13.

`
Figure 2.13: Schematic diagram of the modifications to graphite structure for GNP
preparation [22]
From Figure 2.13, the graphite in GICs can be expanded further through heating at
1000 °C for about 20-30 s; the expansion occurs due to the formation of large
amounts of gases such as H2O and SO2. The resultant expanded GIC is known as
expanded or exfoliated graphite (EG). The heat-expansion process can be

53
accomplished using air or rather a nitrogen atmosphere, since the latter will partially
reduce the formation of graphene oxide (GO) after acid intercalation. In addition to
the heat expansion, microwave irradiation can also be used to obtain further
expansion of GIC as it provides adequate energy that helps to fulfil this purpose.
Afterwards, to obtain final GNPs, an ultrasonication process will be applied for an
adequate period of time in order to bring about further expansion of expanded
graphite [22, 84].

As mentioned earlier, the structure of GNPs comprises sp2 hybridised carbon in the
graphene monolayer; as such, it has extraordinary electrical and thermal properties
[22, 91]. In comparison with other 2D nanofillers such as nano-clay, GNPs possess
low mass density [22, 91]. Moreover, GNPs have an extraordinary combination of
mechanical, electrical and thermal properties, making them an ideal carbon nanofiller
to use for improving the properties of polymer matrices and for use in different
practical applications. Their crucial role in the fabrication of nanocomposites with
excellent electrical, thermal and mechanical properties is to bring about the
homogeneous dispersion of nanofillers within a polymer matrix. By fulfilling this
role, a nanocomposite with enhanced properties will result, which can be used in
different applications such as automobiles, packaging, electromagnetic interference
shielding and electronic devices [92, 93].

2.2.3 Processing Methods of Polymer/GNP Nanocomposites

The dispersion and distribution of GNPs within a polymer matrix is the main issue in
the fabrication of nanocomposites and strongly affects their final properties. As
summarised in the recent reviews by Sengupta et al. [85] and Pang et al. [94], most
nanocomposites are fabricated using in situ polymerisation , solution mixing and
melt compounding. Hence, the following sections will introduce a brief discussion of
the advantages and disadvantages of these techniques.

2.2.3.1 In situ Polymerisation


In situ polymerisation is a successful method that is used to promote the
homogeneous dispersion of carbon nanofillers within polymer matrices. In this
technique, the nanofillers are presence during the polymerisation process of
monomers and/or oligomers, allowing the formation of a strong interaction between

54
reinforcing nanofillers and the matrix phase. As a result, the nanocomposites
produced by this method exhibit a superior improvement in their mechanical
properties and a lower percolation threshold than those produced by solution mixing
and melt compounding, owing to the improved nanofiller dispersion. However, the
main disadvantage of this method is that the processing necessitates high electrical
energy consumption in order to disperse the nanofillers within the matrix during
manufacturing. Hence, this method may not be the best choice for use in the mass
production of polymeric nanocomposites [22, 85, 94].

2.2.3.2 Solution Mixing


In solution mixing, the polymer is dissolved in an appropriate solvent and the filler is
dispersed in the resultant solution. After completing the mixing process, the solvent
is removed and the resultant nanocomposites are usually moulded to the shape of the
composites. The electrical conductivity of nanocomposites made using this method
yield a low percolation threshold due to the good dispersion of fillers. However, the
acceptance of this method in industrial mass production is limited due to the use of
large amounts of solvent, the removal of which may cause environmental pollution
[22, 85, 94].

2.2.3.3 Melt Compounding


Melt compounding is the preferred industrial method that has been used for the mass
production of polymer composites. In this technique, no solvent is used during
composite fabrication, unlike in situ polymerisation and solution mixing, making it
economical and environmentally friendly. Traditional melt-mixing machines such as
extruders and internal mixers can be utilised in this method to manufacture
conductive polymer composites. However, in this method, the composites produced
have a high percolation threshold compared to in situ polymerisation and solution
mixing due to the good quality of nanofillers dispersion leading to form conductive
networks at low weight concentrations [95]. It has been reported that composites
with high percolation threshold experience high melt viscosity and low mechanical
properties. Therefore, reducing the percolation threshold has become the most
pertinent issue in the field of fabricating conductive nanocomposites [22, 85, 94].

To conclude, Albozahid [46] used all three techniques to disperse GNPs within a
TPU-70 HS matrix. The researcher investigated the resultant composites using

55
different characterisation techniques, such as SAXS, WAXD, TGA, DMTA, tensile
testing, and electrical and thermal conductivity tests. The results were confirmed
that better dispersion and lower percolation threshold of GNPs was obtained from
composites manufactured using in situ polymerisation. The researcher revealed that
the composites made by in situ polymerisation become conductive at a percolation
threshold of 5 wt.% GNP loading, in comparison with 10 wt.% GNPs for the
composites made by both solution mixing and melt compounding. Besides, he found
that the in situ polymerisation process can tolerate a loading of 40 wt.% GNPs to
make TPU70-NCs, compared to 10 wt.% GNPs in solution mixing and 20 wt.%
GNPs in melt compounding. It was also demonstrated that the composites made
using in situ polymerisation exhibit a greater improvement in their tensile properties
in comparison with those made using solution mixing and melt compounding due to
the enhancement in the interfacial interaction between GNPs and TPUs structure.
The composites produced by in situ polymerisation exhibit a significant enhancement
in their elastic modulus, while the elongation at break and tensile strength showed a
slight reduction at high loading due to the strain hardening. The thermal conductivity
experienced a dramatic improvement for NCs produced by in situ polymerisation
compared with solution mixing and melt compounding, as illustrated in Figure 2.14.

Figure 2.14: Comparison of thermal conductivity of TPU70/GNP nanocomposites


made by in-situ polymerisation, solution mixing and melt compounding [46]

56
Albozahid [46] also investigated the effect of GNPs and HS content on the properties
of neat TPU and TPU-NCs. The researcher synthesised TPU with different HS
weight percentages (60, 70 and 80) and incorporated different loadings of GNPs
using in situ polymerisation. The results were confirmed his previous results
regarding the effect of HS content on the tensile properties of neat TPU; the TPU-70
HS and TPU70-NCs exhibited moderate tensile strength and better elongation at
break compared to that of neat TPU-60 HS, TPU-80 HS, TPU60/ NCs, and TPU80/
NCs. As a result, in accordance with his study [46] and a study by Al-Obad [47], this
work uses TPU-70 HS as a matrix for NCs and adopts the in situ polymerisation
process to made TPU70-NCs based on different sizes of GNPs and with a loading of
5 wt.%. As suggested by Albozahid [46], the TPU-70 HS incorporated with GNPs
and made by in situ polymerisation reaches a percolation threshold at this loading.

2.2.4 Properties of GNP Nanocomposites

2.2.4.1 Dispersion and Melt Rheology of GNP Nanocomposites


One of the main issues that influences the properties of nanocomposites is their
uniform dispersion and distribution, as mentioned earlier in this chapter. Due to the
Van der Waals forces and dissimilarities in the surface energies of polymer matrices
and nanofillers, nanofillers have a higher affinity towards each other than is the case
for polymer matrices, leading to the formation of nanofiller aggregation [96]. The
aggregation of nanofillers is undesirable in the fabrication of nanocomposites since it
can act as a point of defect in polymer matrices. Indeed, in polymer matrices
reinforced by layered nanofillers such as layered silicates and GNPs, the exfoliation
of stack layers to form smaller individual layers and the stacking of layers to form
aggregation is a prime hurdle that hinders the optimisation of the nanocomposites’
performance. The dispersion qualities of layered nanofillers within polymer matrices
can be illustrated in three common types, as shown in Figure 2.15. The phase-
separated state shown in this figure is obtained when the polymer chains are not able
to penetrate between the graphene-stacked sheets, leading to the formation of
composites with properties similar to those in particulate micro-composites [34, 97,
98]. This occurs because the effective particle size within a polymer matrix can be
greater than 100 nm. The second morphology, known as an intercalated state, is
obtained when the polymer chains can penetrate between the graphene sheets.

57
Therefore, the intercalated morphology of composites can be described as
nanocomposites owing to the effective size of the particles, which measure less than
100 nm [34, 97, 98].

Figure 2.15: Schematic representation of three common descriptions of the


dispersion qualities of layered nanofillers within polymer matrices [97]
The third type in Figure 2.13 represents the exfoliated structure, which is obtained
when full exfoliation and delamination of stacked graphene sheets occurs.
Consequently, a composite with uniform dispersion is achieved and the capacity of
layered nanofillers to efficiently reinforce polymer matrices is optimised [34, 97, 98].

As well as using the SEM, TEM and optical microscopy techniques to investigate the
dispersion of graphene within polymer matrices, melt rheology can also be
considered an effective tool to quantify and show the quality of nanocomposite
dispersion [99, 100]. A number of researchers have undertaken rheological analysis
to investigate the dispersion and quality of nanocomposite structures [25, 101-106].
For instance, Kim and Macosko [106] used melt rheology to study the effect of
consolidating graphite and functionalised graphite sheets (FGS) in a polycarbonate
polymer matrix (PC) after exposure to different annealing times. Both
nanocomposites were prepared using melt compounding. Figure 2.16 illustrates the
dynamic frequency sweeps of PC/G and PC/FGS for different annealing times. It was

58
reported from this figure that the storage modulus (G’) of PC/G and PC/FGS
becomes independent at low frequency at certain concentrations, depending on the
samples’ history and annealing time.

Figure 2.16: Dynamic frequency sweeps of PC/G melts after annealing for (a)
10,000 and (b) 20,000 s, and PC/FGS melts after annealing for (c) 10,000 and (d)
20,000 s at 230 °C [106]
The researchers deduced that the presence of G within the PC matrix without
functionalisation and annealing at 10,000 s had a higher percolation threshold
concentration when the polymer behaved like a solid compared with the PC/FGS
composites. While the PC/G nanocomposites had a percolation threshold of between
3 and 5 wt.% of G, the PC/FGS showed a percolation threshold of between 1 and 1.5
wt.%. This result indicates the improvement in the dispersion and interaction of FGS
with the PC matrix as a result of the functionalisation of G. They also demonstrated
that the extension of the annealing time caused the PC/FGS exhibit a solid-like
response, even at 0.5 wt.%, while the response of the PC/G composites was almost
unchanged event at their percolation concentration.
59
2.2.4.2 Mechanical, Electrical and Thermal Properties of
Nanocomposites
A study was recently conducted by Choi et al. [107] to study the effect of FGS of
different sizes but with similar thickness on the tensile and electrical properties of a
TPU matrix. Their study demonstrated that the percolation threshold of the electrical
conductivity of TPU-NCs decreases with increasing aspect ratio and concentrations
of FGS. It was found that the TPU-NCs incorporated with FGS with a particle size of
2.4 m had a percolation threshold of 1.41 wt.%, while the TPU-NCs blended with
FGS with a particle size of 8.3 m showed a percolation threshold at 0.39 wt.%. The
researchers also reported that the tensile strength and elongation at break reduced
with increasing FGS content, and that this reduction was more evident with FGS
with a large particle size. However, the elastic modulus increased and a significant
improvement was exhibited by the TPU-NCs blended with FGS with a large particle
size. They suggested that the reasons for this disparity in the improvement in tensile
properties could be attributed to the effect of the size of FGS. They theorised that,
during the tensile test, the stress was transferred from the TPU matrix to the FGS,
and this transmission of stress became more effective when the nanofillers had a high
aspect ratio owing to the increase in the interfacial interaction. However, at the same
time, the presence of FGS within the TPU matrix prevented or disrupted the polymer
molecules from reorienting in the direction of elongation, resulting in a reduction in
the elongation and tensile strength. The latter might be ascribed to the weak
interaction between the TPU and FGS, leading to catastrophic fracture, as confirmed
by SEM.

A similar outcome was reported by Gao et al. [108], who investigated the effect of
the size and content of fillers on the morphological, thermal, mechanical and
electrical properties of poly(lactic acid) (PLA)-NCs. They prepared PLA-NCs via
melt compounding and reinforced them with two types of GNPs of large and small
lateral sizes, denoted by GNP-L and GNP-S, respectively. SEM and X-ray
diffraction were used to investigate the dispersion of nanofillers, and both techniques
revealed that the NCs reinforced with GNP-L were well dispersed at weight
percentages  7 wt.%, while the NCs consolidated with GNPs-S had good dispersion
at weight percentages  10 wt.%. The investigation of the tensile properties
confirmed that both nanofillers improved the elastic modulus of PLA/GNP-NCs, but

60
the highest improvement was ascribed to the NCs with GNP-L. This result might be
due to the ability of GNP-L to increase the crystallinity of the PLA-NCs system, as
illustrated in Figure 2.17. Furthermore, the greatest improvement in electrical
conductivity and lower percolation threshold concentration was attributed to the
PLA-NCs consolidated with GNP-L owing to their larger aspect ratio, as has also
been reported in previous studies [107].

Figure 2.17: Young's modulus as a function of crystallinity of PLA (XC (%)) for
PLA/GNP-S and PLA/GNP-L, showing a robust relation between modulus and
crystallinity for the system based on GNP-L [108]
Another study on the effect of filler size on the mechanical and electrical properties
of nanocomposites was carried out by Onyu et al. [23]. The researchers prepared
polypropylene (PP) nanocomposites blended with two different sizes of GNPs via
melt compounding and then prepared samples using injection moulding. The two
GNP types that were used are GNP-M-5 with 5 m average diameter and average
thickness < 10 nm and GNP-H-100 with 200 m average diameter and average
thickness < 15 nm. Both types were studied at two weight percentages, 5 and 7.5
wt.%. It was found that both GNP-M-5 and GNP-H-100 improved the flexural and
electrical properties of PP-NCs. However, the PP-NCs blended with GNP-H-100
showed a greater improvement in the flexural strength than that of GNP-M-5; this
improvement was attributed to the larger size of the former. On the contrary, the
researchers reported a significant improvement in the electrical conductivity of the

61
PP-NCs incorporated with 7.5 wt.% GNP-M-5. They concluded that the greatest
improvement in the electrical conductivity of PP-NCs was attributed to the GNP-M-5
because it had the lowest diameter and thickness but a high surface area, which might
encourage the formation of good dispersion and a conductive network.

A study on the effect of GNP size on the thermal conductivity of nanocomposites has
been carried out by Kim et al. [109]. They used five different types of GNPs that
differed in their lateral size and thickness to fabricate nanocomposites based on
polycarbonate (PC) via melt compounding. The researchers investigated the realistic
size of GNPs and their dispersion within a PC matrix using 3D non-destructive
micro-X-ray CT analysis. This technique revealed that all GNP types had similar
uniform dispersion within the PC matrix regardless of their types, which indicated
the minimal effect of GNP size on their dispersion. Besides, the 3D non-destructive
micro-X-ray CT analysis clearly indicated the effect of the realistic lateral size and
thickness of GNPs on the thermal conductivity of composites. It was found that the
nanocomposites containing GNPs with larger sizes and thickness led to a significant
improvement in thermal conductivity compared to the other types. The researchers
ascribed this improvement to the ability of GNPs with a high aspect ratio to reduce
phonon scattering at the interface between fillers and a matrix and to improve the
ability of the composite to dissipate heat.

2.3 Summary

The previous part of this chapter has given an overview of factors that affect the
formation of phase separation of a TPU structure, followed by its properties. It has
been found that both annealing treatment and hard segment content can help to
increase the microcrystalline hard domain formation and the properties of TPU. It
has also been demonstrated that the TPU with 70 wt.% HS had moderate tensile
properties and a better elongation at break among the lower or higher percentages of
HS. In addition, it has been revealed that the TPU-NCs that give the best tensile,
electrical and thermal properties were those based on the 70 wt.% HS and
consolidated with GNPs via in situ polymerisation. Studies on the effect of nanofiller
size on the properties of nanocomposites have also been discussed in this chapter.
These studies have confirmed that nanofillers with a high surface area can produce

62
NCs with optimal properties due to increasing the interfacial interaction between
nanofillers and the matrix.

2.4 Multiscale Composites

2.4.1 Introduction to Multiscale Composites

Multiscale composites are composites with a hierarchical structure arranged from


nanoscale particles to fibres the size of a micron [110]. Normally, this type of
composite is manufactured by adding nanoscale particles to traditional polymer/CF
composites in order to overcome the weaknesses of the out-of-plane properties and
enhance the properties of the matrix. It is known that the matrix in a composite
system dominates the through-thickness mechanical, electrical and thermal
properties, while the fibre dominates the in-plane properties. Although CF possesses
excellent electrical and thermal conductivity, combining it with a poor conductive
polymeric matrix can reduce the overall conductivity of the resultant composites. It is
also well known that traditional CF-composites are used primarily in aircraft; these
structures have been prone to lightning strikes during take-off and landing, or when
aircraft pass through a storm cloud [111-113]. Subjecting CF-composites to lightning
strikes makes them vulnerable to damage such as fibre breakages, fibre/matrix
debonding, and delamination. This is because composites based on a CF-reinforced
polymeric matrix have poor thermal and electrical through-thickness conductivity,
suggesting that a tremendous amount of energy may not be sufficient to dissipate,
leading to composite damage [111]. To prevent incipient thermal-induced damage in
composite structures and to improve the out-of-plane conductivity, conductive
nanofillers such as graphene and its derivatives can be added to a polymer matrix to
modify its properties [111, 112].

CF-polymeric composites also have poor compressive properties and delamination


resistance, especially those based on a brittle matrix. In fact, the presence of
delamination in these composite structures may lead to catastrophic failure by
provoking a severe reduction in stiffness and strength in the in-plane direction [111,
114, 115]. Delamination can form during subjection to external loadings such as
tension or compression, fatigue cycling, impacts at high or low energy, static bending
or through service or the manufacturing process. A critical cause of delamination is a

63
poor fibre/matrix interface or the brittleness of the matrix [111, 114, 115].
Consequently, this project focuses on using TPU as the matrix for CF composites
due to its excellent toughness and adhesive properties, as mentioned previously.
Using TPU as a matrix might help to overcome the resultant drawbacks of using a
brittle matrix or applying surface treatments to CF. Furthermore, adding nanofillers
to a CF-composite system is another way to improve the through-thickness
mechanical properties like compressive and flexural strength, delamination and
impact resistance. Hence, this project adopts the innovative idea of forming new
multiscale composites (MSCs) based on using TPU70/GNPs-NCs as a matrix
reinforced by CF in order to improve the through-thickness properties, especially
thermal and electrical conductivity. However, the effectiveness of adding nanofillers
to CF-composite systems is strongly influenced by various factors, such as the
concentration and aspect ratio of nanofillers, their dispersion within the polymer
matrix, and the interfacial surface between the fillers, the matrix, NCs and CF [111].
As proved previously, nanofillers with a high surface area result in NCs with
significant improvements in their thermal, electrical and mechanical properties; as a
result of this, this project adopts three different sizes of GNPs. This is in order to
attain the size that best matches the TPU70/CF composites to give a good
improvement in thermal, electrical and mechanical properties.

2.4.2 Carbon-Fibre Reinforcement


Carbon fibres (CFs) can refer to fibres that contain at least 92 wt.% carbon in their
compositions [116, 117]. Commercially, CFs can be found in three basic forms: long
and continuous tow, chopped (6-50 mm), and milled (30-3000 m) [34]. For
decades, carbon fibres have contributed to many industries such as the chemical
industry, nuclear energy, missiles, automobiles, aerospace and the composite
materials industry, due to their intrinsic properties. CFs have many advantages,
which include a high tensile-weight ratio and a high modulus-weight ratio,
dimensional stability which relates to their low coefficient of thermal expansion
[34], high thermal conductivity (higher than the thermal conductivity of copper),
high fatigue strength and excellent vibration damping, high electrical conductivity,
and good wear and friction resistance. Moreover, carbon fibres do not experience
stress corrosion or stress rupture failure at ambient temperature compared with other

64
fibres such as glass and organic polymer fibres [34, 118, 119]. CF can be classified
into three categories: high-strength, intermediate-modulus and high-modulus. Table
2.1 illustrates these three types of CF with their representative properties [120].
Table 2.1: Carbon fibre properties [120]

Property High-strength Intermediate- High-modulus


(HS) modulus (IM) (HM)
Diameter (m) 6-8 6-9 7-9

Density (g/cm3) 1.7-1.8 1.74 1.85-1.96

Tensile strength (MPa) 3000-5600 4800 2400-3000

Elongation at break (%) 1.0-1.8 2.0 0.38-0.5

Young’s modulus (GPa) 235-295 296 345-520

Specific strength (106 cm) 17.5-32.7 28.2 15.7

Specific modulus (106 cm) 1370-1720 1740 1850-2790

Coefficient of thermal
expansion (10-6/K)
Axial -0.5 - -1.2
Radial 7 - 12

In general, the mechanical properties of carbon fibres are strongly influenced by the
compositions of their precursors, which are commonly polyacrylonitrile (PAN),
rayon (cellulose) and pitch, in addition to the temperature and time of the production
process. The manufacturing process of CF can be divided into five steps: fibre
precursor production, oxidisation, carbonisation, graphitisation and surface treatment
or sizing treatment [117]. In reality, higher processing temperatures produce CF with
a greater degree of crystalline orientation in the fibre axis, and a higher fibre modulus
[120].
Continuous CF is normally produced as fibre tow that is wound onto a reel. Each tow
consists of a bundle which comprises thousands of continuous CF filaments, usually
known as 3, 6 and 12 k. These bundles are held together and coated or sized using
organic materials to protect them from the environment. CF can be produced in
different configurations ranging from unidirectional CF to multi-axial fabric in order
to achieve the desirable properties for its various applications. These fibre
configurations are produced using different techniques to obtain stitched, woven,
braided or multi-axial knitted fabrics; moreover, CF can be arranged as 2D or 3D
fabrics.

65
2.4.3 Manufacturing of Multiscale Composites

The manufacturing of multiscale or hierarchical composites depends firstly on


dispersing the nanofillers within a polymer matrix using a traditional approach such
as shear mixing or ultra-sonication [114]. The following step is to infuse the
dispersed matrix resin into the fibre using vacuum-assisted resin transfer moulding
(VARTM) or resin transfer moulding (RTM). These approaches are compatible with
industrial processes [114]. However, they are limited to fabricating MSCs based on
thermoplastic (TP) or thermoplastic NCs due to the high viscosity of the matrix,
which might lead to the formation of high void percentages owing to the lack of fibre
wetting [114, 121]. Therefore, a hot-press or compression-moulding process is a
suitable process for forming MSCs based on matrices from TP and its NCs [9, 19,
122-124]. For example, Díez-Pascual et al. [9, 123] used the hot-press process to
manufacture MSCs comprised of polyetheretherketone (PEEK) as a matrix,
reinforced with SWCNTs and 5-harness (H) glass-fibre (GF) fabric. Firstly, the
researchers used melt compounding to form PEEK-NCs using two types of
SWCNTs: neat SWCNTs and one wrapped in a compatibiliser made from
polyetherethersulfone (PEES). Following this, a thin film with a thickness of 500 m
was prepared via a hot-press process. Then four alternating layers each of 5H-GF
fabric and PEEK/SWCNT films were formed using a hand lay-up method and
compressed inside a brass frame to maintain the composite dimensions at 380 °C for
30 minutes. The resultant MSCs had a fibre volume fraction of around 48% and void
content of less than 3%.

2.4.4 Effect of Nanofillers on the Properties of MSCs

Substantial research has been dedicated to studying the effect of adding nanofillers to
enhance the through-thickness properties of MSCs [9, 10, 17, 20, 123, 125-130].
Since these materials are used in structural applications, studying the effect of
nanofillers on the mechanical properties dominated by the matrix, such as flexural,
impact resistance, and delamination resistance, are crucial. However, using TPU or
TP/GNPs-NCs as a matrix and reinforcing it with CF fabric has very rarely been
attempted. Therefore, other research that deals with different types of matrices, fibres
and nanofillers will be used as references to describe and dispute the effect of
nanofillers on the mechanical, electrical and thermal properties of MSCs.

66
2.4.4.1 Tensile and Flexural Properties
The previous work mentioned by Díez-Pascual et al. [9] studied the effect of adding
neat SWCNTs and SWCNTs wrapped with PEES compatibiliser on the tensile and
flexural properties of PEEK/GF composites. The researchers found that the
PEEK/GF composites consolidated with neat and wrapped SWCNTs exhibited a
significant improvement in their flexural properties, while the tensile properties
improved slightly in comparison with neat PEEK/GF composites. The greatest
improvement in the flexural and tensile modulus of MSCs was ascribed to the MSCs
incorporated with wrapped SWCNTs, owing to the improvement in the interfacial
interaction between the SWCNTs and the PEEK matrix which occurred upon
modification with PEES compatibilisers.

More research on studying the effect of nanofiller dispersion on the flexural and
tensile properties of CF-composite was performed by Kim et al. [130]. The
researchers dispersed 3 wt.% CNT within epoxy resin via an ultra-sonication
technique using different sonication times (10 minutes and 3 hours). The VARTM
technique was then used to infuse a dispersed resin into CF fabric to produce MSCs.
It was found that adding 3 wt.% CNT to the epoxy/CF system had a slight effect on
the tensile properties, while a moderate improvement in the flexural properties was
recorded in comparison with controlled epoxy/CF composites. The highest
improvement was attributed to the MSCs with GNPs sonicated for 3 hours compared
with those sonicated for 10 minutes. Their findings were confirmed by SEM in order
to investigate the fracture surface of the MSCs. While the SEM images confirmed
poor dispersion and weak interfacial adhesion of CNTs dispersed for 10 minutes,
which was the reason for the slight improvement in both the flexural and tensile
properties, good dispersion was observed for the MSCs with CNTs dispersed for 3
hours.

Another recent investigation was accomplished by Ávila et al. [131] to study the
effect of incorporating GNPs into epoxy/CF composites on the bending stiffness and
strength. The researchers used sonication and a shear-mixing process to disperse
GNPs within epoxy resin by following the method that was described in the previous
nanocomposite studies by their group [132]. They observed a significant
improvement in the bending stiffness and strength of epoxy/GNPs/CF composites
with 0.5 wt.% GNPs compared to controlled epoxy/CF, while this improvement
67
fluctuated at high weight percentages of GNPs. It was suggested that the variation in
this improvement of bending strength could be attributed to the formation of GNP
agglomeration at high concentrations, resulting in various changes in the failure
mechanisms. The SEM technique proved that the controlled epoxy/CF composites
experienced an intralaminar failure mode, while MSCs with high concentrations
showed a mixed failure mode consisting of inter- and intralaminar failure that
appeared as a zigzag pattern, as shown in Figure 2.18.

(a) (b)

Figure 2.18: SEM images to describe the failure modes of (a) controlled epoxy/CF
composites and (b) epoxy/CF incorporating 1 wt.% GNPs [131]

2.4.4.2 Interlaminar Shear Strength (ILSS)


Junid [10] studied the effect of adding GNPs to epoxy/CF composites on their
mechanical properties. The researcher treated GNPs with 3-
aminopropyltriethoxysilane (APTS) donated by A-GNPs and used it to modify
epoxy/CF composites via dispersing them within an epoxy matrix using a sonication
technique. A vacuum-infusion process was used to impregnate the CF fabric with the
dispersion resin in order to produce MSCs. His study revealed that all of the weight
percentages of A-GNPs that were used to manufacture MSCs caused a reduction in
their ILSS value from that of neat epoxy/CF composite. It was reported that the
highest reduction was attributed to the high A-GNP concentration, which was 6
wt.%. He interpreted this reduction as due to the poor adhesion between the NC
matrix and CF, as confirmed by SEM images and to the increase in the viscosity of
the matrix due to the incorporation with A-GNPs, which could also affect the fibre
impregnation. Furthermore, the presence of A-GNP agglomeration may also be a
reason for the reduction in the ILSS value, as they act as stress-concentration points.

68
Rahman et al. [133] investigated the behaviour of epoxy/GF incorporating amino-
functionalised MWCNTs (NH2-MWCNT) after ILSS testing. The researchers used
four different weight percentages of NH2-MWCNTs (0.1, 0.2, 0.3 and 0.4) and
dispersed them within epoxy resin using sonication. The mixture of epoxy
nanocomposites was then used to produce MSCs using a hand lay-up method
followed by a hot-press method in order to attain good GF wetting and homogeneous
distribution of the epoxy matrix within the GF fabric. It was observed that the ILSS
value of MSCs increased with increasing NH2-MWCNT concentration. The
maximum improvement was reported for the epoxy/GF containing 0.3 wt.% NH2-
MWCNT, while the ILSS value decreased for MSCs with 0.4 wt.% of NH2-
MWCNT, as illustrated in Figure 2.19.

Figure 2.19: Effect of NH2-MWCNT concentration on the ILSS values of epoxy/GF


composites [133]
The researchers ascribed this behaviour to the good interfacial interaction between
the functionalised MWCNT and the epoxy matrix, leading to an increase in the load
transferred between them. However, at high concentrations of MWCNTs, the high
surface area of MWCNTs led to an increased attraction between the MWCNT
particles themselves, leading to agglomeration. It was also found that the high
content of MWCNTs caused an increase in the viscosity of the epoxy/NC matrix,
causing a void or bubbles to form and poor fibre-wetting as a result; this
phenomenon may lead to reduced interfacial adhesion between NC matrices and CF.
69
Another study was accomplished by Liu et al. [134]. The researchers studied the
effect of poly(phthalazinone ether ketone) (PPEK) as an agent containing 1 wt.%
CNTs that was used to size a unidirectional CF tow on the ILSS of PPEK/CF
composites. They used a prepreg technique to form PPEK/CF composites and
PPEK/CNT/CF composites. Their results revealed that the MSCs reinforced with CF
with a mixture of the sizing agent PPEK/CNTs showed an improvement in ILSS of
115% compared with neat PPEK/CF composites. It was suggested that the reason for
this improvement was attributed to the good interfacial adhesion between the PPEK
matrix and the sized CF owing to the increased surface area of CF, leading to
enhanced stress transfer from the matrix to the fibre and vice versa.

2.4.4.3 Interlaminar Fracture Toughness (ILFT)


One of the most critical mechanical properties that is dominated by matrices of
composite materials is the ILFT. ILFT is an important property of laminated
composites that can be determined via the propagation of pre-cracks and expressed in
terms of the critical strain energy release rate (GC). GC can be defined as the amount
of energy consumed by materials in order to grow a delamination through a unit area
within composites [135]. In an ILFT test there are three different modes of loading
test, as shown in Figure 2.20 [136]. The first one is denoted “mode I” and represents
the crack opening; the second one is the mode II or in-plane shear; and the last one is
mode III or mixed mode, which represents the anti-plane shear. Mode I and mode II
are the most common loading modes in ILFT tests. During these tests, GC can be
measured in two stages. The first one is called the initiation GC and this relates to the
propagation of the crack at the end of the pre-crack film; the other stage occurs when
the crack propagates within a sample.

Figure 2.20: Schematic representation of three loading modes during interlaminar


fracture toughness test [136]

70
Because of the crucial effect of delamination resulting from low interlaminar fracture
toughness of CF composites’ matrices, this study adopted the mode-I ILFT test to
investigate the effect of GNP loading on the ILFT of TPU70/CF laminates. At the
time of writing, no studies of the ILFT of MSCs consisting of TPU, GNPs and CF
were found; as such, different MSC systems are adopted to discuss and describe the
effect of nanofillers on the mode-I ILFT of MSCs. Wang et al. [137] investigated the
effect of modifying an epoxy matrix by adding neat and functionalised stacked-cup
carbon nanofibers (SCCNFs) on the mode-I and mode-II ILFT of epoxy/CF
laminates. The researchers used the sonication process to disperse neat and
functionalised SCCNFs in an epoxy matrix; the dispersed resin was then used to
produce MSCs made from epoxy/SCCNFs/CF using a hand lay-up technique. Their
results confirmed that the presence of 0.5 wt.% and 1 wt.% of functionalised
SCCNFs within the epoxy matrix enhanced the mode-I and mode-II ILFT by 13.6%
and 45.3% respectively from that of pristine epoxy/CF laminates. It was found that
the results of the mode-I ILFT test exhibited less improvement than the mode-II
results. This outcome was attributed to the nature of both tests: while mode-I is the
tensile opening mode, which is controlled by the strength of the interface between
SCCNFs and epoxy, mode-II is the in-plane shear, which is dominated by the
properties of the matrix [137]. SEM confirmed the enhancement in ILFT of MSCs in
both modes due to the role that SCCNFs play in hindering the crack propagation and
increasing energy dissipation. For example, Figure 2.21 shows the failure
mechanisms of mode-I ILFT, which depends on the alignment of the SCCNFs
regarding the crack propagation.

Figure 2.21: SEM images of failure mode after mode-I ILFT test of epoxy/0.5 wt.%
SCCNFs/CF laminates. (a) Peel off and pull-out of SCCNFs, and b) unravelling of
SCCNFs [137]

71
Therefore, the effect of SCCNFs within MSCs can be to bridge cracks if they are
perpendicular to the direction of crack propagation (see Figure 2.21 a), pulled out
when the tensile load increases, depending on the SCCNFs-epoxy interface, and
finally unravelling graphene layers of CNFs to a longer distance from the opening
crack distance (Figure 2.14 b).

Similar research was performed by Yokozeki et al. [138], who studied the effect of
modifying an epoxy matrix by dispersing 5 wt.% of stacked-cup CNTs (SCCNTs) on
the mechanical properties of unidirectional epoxy/CF laminates. Their study on the
mode-I and mode-II ILFT behaviour of epoxy/CF confirmed an improvement in its
values after adding SCCNTs. During the mode-I test, they visually observed a small
amount of fibre bridging, which may be the reason for this improvement in the ILFT
value in the presence of SCCNTs. Meanwhile, the SEM images of the fracture
surface after mode-I and mode-II tests confirmed an increase in the roughness of the
fracture surface in comparison to that of neat epoxy/CF due to the presence of
SCCNTs. It was suggested that the dispersion of SCCNTs assists with extending
crack deflection, which might help to increase the ILFT value.

A further study was accomplished by Borowski et al. [127], in which the researchers
investigated the effect on the mode-I ILFT value of incorporating carboxyl
functionalised MWCNTs (COOH-MWCNTs) within epoxy/CF laminates. Their
results showed that the addition of 0.5%, 1% and 1.5% weight percentages of
COOH-MWCNTs caused an increase in the ILFT value of MSCs by 25%, 20% and
17% respectively, in comparison with neat epoxy/CF. It was supposed and confirmed
using an optical microscopic technique that the presence of MWCNTs hinders the
crack propagation, which causes an increase in the consumption energy due to the
occurrence of crack branching. However, high concentrations of MWCNTs caused
an increase in the matrix viscosity, which affects the impregnation of CF, leading to
reduced enhancement of the ILFT value.

2.4.4.4 Impact-damage Resistance and Damage Tolerance


Since CF composites are used in various structural applications, such as the
aerospace, automotive and marine industries, it is essential to address one of their
major limitations during service. Impact damage to CF-composite structures can be a
major issue that limits their application and service life. These types of damage are

72
generated when CF composites are subject to a low-energy impact, such as dropping
a maintenance tool or colliding with runway debris and is called barely visible
impact damage (BVID) [1, 7, 8], indicating the difficulty of detecting them by visual
inspection. Moreover, the presence of BVID can have a significant effect on
reducing the compressive strength and durability of composites. Typical impact
damage during subjection to low-energy impact comprises matrix-cracking
delamination and fibre breakage; the latter is only supposed to occur during
subjecting to high-energy impact [1, 122, 139]. Figure 2.22 illustrates the typical
impact-damage modes of laminated composites upon subjection to low-energy
impact.

Figure 2.22: Schematic representing typical damage modes of low-energy impact of


laminated composites [6]
In this figure, the damage modes within the impactor area consist of buckling of
delaminated plies due to the compressive load that is applied by the impactor. The
matrix cracks formed inside the composite can be created at the top as a result of
contact stress, in the middle due to the shear, and on the bottom surface due to
bending [1, 6].

In order to assess the impact-damage tolerance of laminate composites, compression


after impact tests can be applied. As mentioned earlier in this section, the generation
of delamination within laminated composites can lead to the reduction of the
compressive and residual compressive strength, which stimulates premature buckling
failure [1, 7]. The residual compressive strength is normally calculated as the ratio of
the compressive strength after being subjected to impact to the compressive strength
without impact. Recently, due to progress in the application of nanofillers, especially

73
of graphene and its derivatives, many researchers have tried to use them as an
additive to laminated composite systems [122, 138, 140-144]. This is in order to
improve the matrix toughness, leading to improvements in the impact-damage
resistance and impact-damage tolerance of the composites [122, 138, 140-145].

Ulus et al. [145] used a hybrid nanoparticle based on MWCNTs and boron nitride
nanoplatelets (BNNP) to modify epoxy/CF composites. They manufactured four
types of composites (neat epoxy/CF, epoxy/MWCNTs/CF, epoxy/BNNP/CF and
epoxy/hybrid (BNNP/MWCNTs)/CF) using an infusion technique to impregnate 10
plies of woven CF. The researchers investigated the effect of these nanofillers and
hybrid nanofillers on the impact resistance of the resultant composites by subjecting
composites to low-velocity impact tests at 5, 10 and 15 J. It was observed that the
addition of BNNP and BNNP/MWCNTs could improve the impact resistance more
significantly than MWCNTs via increasing the maximum impact load that
composites can resist during the test. This is because of the ability of MWCNTs to
restack due to their Van der Waals forces leading to form agglomeration which
works as a stress concentration points. For example, at 15 J the addition of BNNP
and BNNP/MWCNTs improved the maximum impact load of MSCs by 25 % and 50
% respectively compared to conventional epoxy/CF, while the MWCNTs rose the
maximum impact load by 12 %. Therefore, both MSCs incorporated with BNNP and
hybrid nanofillers experienced lower matrix cracking and delamination in
comparison with neat epoxy/CF and epoxy /MWCNTs/CF composites. In general,
the modified epoxy/CF showed a significant improvement in impact-damage modes
controlled by the matrix, such as matrix cracking and delamination, in comparison
with those of neat composites, indicating the success of the modifications using
nanofillers.

Kostopoulos et al. [122] investigated the effect of modifying epoxy/CF composite by


dispersing 0.5 wt.% MWCNTs in an epoxy matrix on the impact and post-impact
behaviour of neat epoxy/CF composites. The researchers used a high-shear device to
disperse the MWCNTs in an epoxy resin that was later used to manufacture MSCs
via a vacuum-infusion process. The MSCs and neat epoxy/CF were subjected to low-
energy impact at 2 J, 8 J, 12 J, 16 J and 20 J energy levels, and were then subjected
to compression tests to assess the damage tolerance. Their results demonstrated that
the MSCs clearly experienced a smaller delamination area than that of the neat

74
epoxy/CF, while the absorbed energy showed a slight increase at high impact energy,
as illustrated in Figure 2.23 a and b respectively.

(a)

(b)

Figure 2.23: The effect of incorporating MWCNTs within epoxy/CF composites on


the a) delamination area against impact energy levels and b) the percentage of
absorbed energy against energy levels for MSCs and neat composites [122]

75
Meanwhile, the CAI strength decreased with increasing impact energy levels, and the
highest reduction was exhibited at higher impact energy levels. However, the MSCs
exhibited a higher CAI strength value than that of neat composites at all impact
energy levels. They reported that the inclusion of MWCNTs led to an increase in
CAI strength by approximately 12-15% and in the CAI modulus by 15%. These
results may indicate the success of MWCNTs in toughening the epoxy matrix and
improving the CAI strength of MSCs.
Eskizeybek et al. [144] modified an epoxy matrix by using three different weight
percentages of CaCO3 nanoparticles and used the dispersed matrix to manufacture
MSCs. They used the VARIM technique to impregnate CF with the modified epoxy
matrix; the resultant MSCs were used to study the effect of CaCO3 nanoparticles on
their impact-damage resistance behaviour compared to neat epoxy/CF. It was found
that the MSCs with a high loading of CaCO3 showed a higher increase in absorbed
energy and peak deflection, while the lower percentages ( 2 wt.%) showed a
significant reduction in the absorbed energy compared to the neat composites. These
results demonstrated that the CaCO3 particles succeeded in toughening the matrix by
improving their ability to resist the creation and development of matrix cracking and
delamination, suggesting a reduction in the absorbed energy. However, the high
loading (3 wt.%) of CaCO3 led to the creation of agglomeration that acted as stress
concentration, which then stimulated the creation of delamination and matrix cracks,
as was confirmed by SEM.

2.4.4.5 Electrical and Thermal Properties


As mentioned previously in this chapter, the addition of graphene and its derivatives
to CF composites can help to improve their electrical and thermal properties.
Enhancements in the electrical and thermal conductivity of CF composites due to the
addition of conductive nanofillers have been reported by many researchers [115, 123,
129, 146-148]. For example, Li et al. [148] modified epoxy/CF composites using two
loadings of GNPs (2 wt.% and 5 wt.%). They consolidated an epoxy matrix with
GNPs using a three-roll mill (TRM) process; the dispersed resin was used to
impregnate the CF using a vacuum infusion process. The resultant MSCs were used
to study the effect of GNPs on the electrical and thermal properties in the in-plane
and out-of-plane directions. It was demonstrated that the addition of 5 wt.% GNP
caused a slight enhancement in in-plane electrical and thermal conductivity, while

76
the out-of-plane conductivities increased by 50% in comparison with neat epoxy/CF
composites. The researchers attributed the improvement in electrical conductivity to
the ability of GNPs to form a 3D conductive network that helps to transmit electrons
between the plies in the out-of-plane direction. However, the enhancement in out-of-
plane thermal conductivity was attributed to the ability of GNPs to reduce the
mismatch in thermal conductivity between the matrix and the CF, where the latter
controls phonon scattering. This is because the flat shape and high surface area of
GNPs help to reduce the phonon scattering caused by the matrix via increasing the
areal contact between neighbouring platelets within the matrix itself.

Qin et al. [115] modified epoxy/CF composites by coating the CF by immersing


them in a homogeneous GNP suspension. Following this process, and after placing
the coated CF within a drying tower, the coated CF was used to form MSCs via
impregnation with epoxy. The epoxy/coated CF composites were then used to study
the effect of GNPs on the in-plane and out-of-plane electrical conductivity. Their
results confirmed that the MSCs with coated CF increased the out-of-plane electrical
conductivity by 165% compared to neat epoxy/CF composites. The researchers
attributed this result to the ability of GNPs to create a contact pathway in the
interface between the CF and the epoxy matrix. It was suggested that the migration
of GNPs away from the CF surface might occur and, owing to the compaction of CF
through the manufacturing process, the thickness of the coating layer on the surface
of the CF might increase from that of the original one before fabrication.

A recent study has been published by Imran and Shivakumar [147], in which the
researchers studied the effect of dispersing 1 wt.% of GNPs within epoxy/CF
composites on their electrical and thermal properties. They used a TRM process to
disperse the GNPs within epoxy resin. Following that, the dispersed resin was used
to form MSCs via a hand lay-up technique of eight plies of CF, followed by vacuum-
bagging compression. The eight plies of CF were used to investigate the electrical
conductivity, while the measurement of thermal conductivity was carried out for
composites of two plies of CF. Their results confirmed that the presence of GNPs
increased the electrical conductivity to double the value of neat epoxy/CF
composites, and the thermal conductivity exhibited a slight increase from 0.68 to
0.72 W/m.K. They attributed the improvement in electrical conductivity to the
GNPs’ role in forming an extra conductive pathway between the MSC laminates.

77
However, it was reported that the thermal conductivity of MSCs exhibited a slight
improvement of 6% in comparison with that of neat composites, suggesting that the
differences in CF volume fraction (Vf) were more significant than the incorporation
of GNPs: when the Vf of neat composites was 57%, the Vf of MSCs was 51%.

2.5 Summary

This chapter has given an overview of the synthesis of TPU polymers and has
concisely explained the factors that affect their phase separation and their properties.
It is well known that the most important element of a TPU structure is the formation
of phase separation due to the incompatibility between SS and HS. Hence, it has
been necessary to understand the factors that can increase or reduce the formation of
phase separation, from the ratio of SS and HS to the annealing treatment. Normally,
annealing treatment is used to improve the crystallinity and then the properties of
TPU by stimulating the HS to reorder and increase the degree of formation of phase
separation. With regard to previous studies of the effect of HS on mechanical
properties, it was found that TPU based on 70% HS can give better tensile properties
than higher percentages of 80% or 90 %, as the latter cause the TPU to be stiffer and
failed in a short elongation at break during tensile tests. As a result, this project
adopts TPU with 70 wt. % HS as a matrix to reinforce with different sizes of GNPs
in order to produce nanocomposites with high performance in electrical, thermal and
mechanical properties.

GNPs, on the other hand, have excellent electrical, thermal and mechanical
properties and low production costs when compared to the cost of CNT production.
Within the aim of this study in incorporating GNPs within a TPU structure, both the
dispersion and distribution of nanofillers have a great effect on the resultant NC
properties. Therefore, this project uses in-situ polymerisation to consolidate GNPs
within TPU, as this was reported previously as being the best method to form
TPU/GNPs-NCs with better electrical, thermal and mechanical properties. Different
research studies have been reviewed to study the effect of incorporating GNPs or
other graphene derivatives within thermoplastic matrices on their rheology behaviour
and mechanical, thermal and electrical properties. Furthermore, this chapter has
introduced the concept of studying the effect of surface area and aspect ratio of
nanofillers on the properties of resultant nanocomposites.

78
Due to the low toughness and damage tolerance of epoxy/CF composites, this
research uses TPU-70 wt.% HS instead of epoxy to form CF composites due to its
excellent toughness, ductility and adhesive properties. This research is novel, in that
no other studies have been found that have made a composite based on TPU as a
matrix and CF fabric as reinforcement. As outlined previously, the low electrical and
thermal properties of a TPU matrix can lead to a reduction in the electrical and
thermal properties of resultant TPU/CF composites, especially in the out-of-plane
direction. In addition, it was reported that the matrix is responsible for most of the
electrical, thermal and mechanical properties of composites in the through-thickness
direction. Therefore, many researchers have tended to improve the matrix through-
thickness properties via consolidation with different conductive nanofillers such as
graphene, GNPs, G and CNTs.

Work with MSCs is still new; as such, no research has been found which uses
TPU/GNPs-NCs with fabric CF, suggesting that other MSCs can be used to describe
and discuss the effect of conductive nanofillers on the properties of MSCs. It has
been reported that nanofillers have the ability to improve the mechanical properties
of MSCs, such as tensile and flexural modulus, ILSS, mode-I interlaminar fracture
toughness, impact-damage resistance and damage tolerance.

Hence, in this work the TPU-70 HS is modified using different sizes of GNPs which
are 5, 15 and 25 m in average particle diameter. Those TPU70-NCs are prepared
via in situ polymerisation which recommended as a best technique to produce a
uniform dispersion of GNPs within TPU-70 HS matrix than solution mixing and melt
compounding. Then the neat TPU-70 HS and TPU70-NCs are used to form MSCs
via compression moulding process in order to study the effect of GNPs sizes on the
thermal, electrical and mechanical properties of TPU70/CF laminates. The candidate
MSCs which give better thermal, electrical and mechanical properties can then be
used to study the effect of GNPs content on the delamination resistance, impact-
damage resistance and damage tolerance of the TPU/CF composites.

79
Chapter 3: Experimental Methods

3.1 Introduction

In this chapter, the description of the chemical raw materials that were used to
synthesise TPU-70 HS and its nanocomposite will be explained in detail. In addition,
a description of the conductive raw materials that were used to manufacture
TPU70/CF laminate and multiscale composites (MSCs), consisting of graphene
nanoplatelets (GNPs) and carbon fibre (CF), will be given. The second part of this
chapter will describe the manufacturing process that was utilised to prepare the
nanocomposite and MSCs samples. The cutting process that was used to prepare the
MSCs specimens for different mechanical tests and the SEM technique will also be
explained. The other parts of this chapter will explain and describe all the
characterisation techniques that were carried out on the GNPs fillers, nanocomposites
and MSCs specimens. At the end of this chapter, an explanation of the mechanical
tests that were used during the present work will be given, as well as their
experimental conditions and theoretical analyses.

3.2 Materials

3.2.1 Raw Materials of TPU-70 HS Synthesis

3.2.1.1 Isocyanate
MDI (4,4-methylenebis phenyl isocyanate), supplied in solid form by Sigma-Aldrich
UK, was one of the chemical materials used to produce TPUs. The 4,4-MDI
chemical structure and its properties are illustrated in Figure 3.1 and Table 3.1 [149].

Table 3.1: Properties of 4,4’-methylenebis phenyl isocyanate [149]

Properties of MDI
̅ w (g/mol)
Average Molecular Weight, M 250.25
Melting Point (°C) 42-45 (lit.)
Boiling Point (°C) 200 at 7 hPa (lit.)
Purity (%) 98
Density (g/mL) 1.18 at 25 °C
Storage Temp. (°C) -20

80
Figure 3.1: Chemical structure of 4,4’-methylenebis phenyl isocyanate [149]

3.2.1.2 Polyol
The polyol, which was also supplied by Sigma-Aldrich UK, and which was used as a
soft segment during this work, was based on a polyethylene glycol-block-
polypropylene-glycol-block-polyethylene-glycol structure (PEG-PPG-PEG) with
functionality 2.0. The chemical structure and properties of a PEG-PPG-PEG polyol
is illustrated in Figure 3.2 and Table 3.2 [149]..

Figure 3.2: Chemical structure of PEG-PPG-PEG polyol [149]


Table 3.2: Properties of PEG-PPG-PEG polyol [149]

Properties of Polyol
̅ w (g/mol)
Weight-average molecular weight, M 2000
Boiling Point (°C) > 149 °C
Density (g/mL) 1.006 at 25 °C
Purity (%) Not Available

3.2.1.3 Chain Extender


The 1,5-pentanediol chain extender, which was also supplied by Sigma-Aldrich, was
used to form hard segments with 4,4-MDI in the final stage of TPU synthesis. The
chain extender chemical structure and its properties are illustrated in Figure 3.3 and
Table 3.3.

Figure 3.3: Chemical structure of chain extender (1,5-pentanediol) [149]

81
Table 3.3: Properties of chain extender (1,5-pentanediol) [149]

Properties of Chain Extender (1,5-pentanediol)


Boiling Point (°C) 242
Density (g/cm3) at 25 (°C) 0.994
̅ w (g/mol)
Weight-average molecular weight, M 104.15
Purity (%)  97.0

3.2.1.4 Catalyst
1,4-Diazabicyclo[2.2.2]octane (DABCO), which was also supplied by Sigma-Aldrich
in solid form, was used in the synthesis of TPUs as a catalyst, as is commonly the
case. The chemical structure and properties of DABCO are displayed in Figure 3.4
and Table 3.4 [149].

Figure 3.4: Chemical structure of 1,4-Diazabicyclo[2.2.2]octane [149]


Table 3.4: Properties of 1,4-Diazabicyclo[2.2.2]octane [149]

Properties of 1,4-Diazabicyclo[2.2.2]octane
Weight-average molecular weight, M ̅ w (g/mol) 112.17
Melting Point (°C) 156-159 (lit.)
Vapour Pressure mmHg (50 °C) 2.9
Density (g/mL) 1.02 at 25 °C (lit.)
Purity (%)  99.0

3.2.1.5 Solvent
N, N-Dimethylacetamide (DMAc) was used as a solvent during the synthesis of
TPUs, and was again supplied by Sigma-Aldrich. The chemical structure of DMAc
and its properties are presented in Figure 3.5 and Table 3.5.

Figure 3.5: Chemical structure of N,N-Dimethylacetamide [149]

82
Table 3.5: Properties of N,N-Dimethylacetamide [149]

Properties of N,N-Dimethylacetamide (DMAc)


Density (g/mL) 0.937 at 25 °C (lit.)
Boiling Point (°C) 164-166 at 760 mm Hg

Melting Point (°C) -20


̅ w (g/mol)
Average Molecular Weight, M 87.12
Purity (%)  99.75

3.2.2 Conductive Raw Materials

3.2.2.1 Graphene Nanoplatelets (GNPs)


In this work, xGnP® graphene nanoplatelets (GNPs), grade M, were purchased from
XG Sciences, US, and used as a filler. GNPs are distinctive nanoparticles consisting
of short stacks of graphene sheets with a platelet shape. Grade-M GNPs are available
in three different average particle diameters, M5, M15, and M25, with average
diameters of 5, 15, and 25 m respectively. These grade-M nanoparticles possess an
average thickness of approximately 6-8 nm and have a typical surface area of
between 120 and 150 m2/g. Their other properties are tabulated in Table 3.6 [150].

Table 3.6: Typical properties of grade-M GNPs [150]

Properties Parallel to surface Perpendicular


to surface
Density (g/cm3) 2.2
Thermal conductivity (W/m.K) 3,000 6
Thermal Expansion (CTE) (m/m/K) 4-6  10-6 0.5-1.0  10-6
Tensile Modulus (GPa) 1,000 N/A
Tensile Strength (GPa) 5 N/A
Electrical Conductivity (S/m) 107 102

3.2.2.2 Carbon Fibre


The carbon fibre (CF) used in this project was purchased from Sigmatex UK, and
consisted of a five-harness satin (5HS) woven fabric with a fabric areal weight
(FAW) of 380 g/m2. The properties of the 5HS carbon fibre fabric are listed in Table
3.7 [151].

83
Table 3.7: Properties of 5HS carbon fibre fabric [151]

Name Sigmatex-5HS
Fibre Description 380GSM/5HS/T400HB 6K 40D/1295mm
Tensile Strength (MPa) - Nominal 4410
Tensile Modulus (GPa) - Nominal 250
Density (g/cm3) - Nominal 1.8
Filaments per Tow 6000.0
Width (mm) 1295

3.3 Synthesis of Unfilled TPU

3.3.1 Preparation of Synthesis Apparatus

All the apparatus for the polymerisation of TPU-70 HS and TPU70/nanocomposites


(NCs), such as glassware and accessories, was dried at 70°C in an oven for one night

before being used. This process was carried out in order to remove any moisture
which could affect the synthesis process. After that, the equipment was stored in an
oven until they were used.

3.3.2 Preparation of Chemical Substance for Synthesis

Before commencing synthesis, it was crucial to dry the chain extender and polyol in
order to prevent any side reactions that could occur between the isocyanate group in
the MDI structure and the hydroxyl group due to moisture. These side reactions
could affect the isocyanate groups by hydrolysing them, leading to the formation of
carbon dioxide and amine. Therefore, both the chain extender and polyol were dried
at 100 °C in a vacuum oven overnight. Besides, the molecular sieves, which were in
turn used to prevent any moisture from contaminating the dried polyol and chain
extender, were dried in an oven at 300 °C for three days. The dried molecular sieves
were then added to both the dried chain extender and polyol and stored in sealed
glass jars until synthesis.
The DMAc and DABCO, which were used in this work, were stored in a dry place in
supplied sealed containers. Meanwhile, the 4,4-MDI was preserved as a solid in a
refrigerator in sealed jars at -20 °C. Finally, after each synthesis, nitrogen gas was

injected into the bottles or jars of DABCO, DMAc and MDI in order to prevent any
moisture from entering and contaminating the materials. The bottles or jars of

84
DABCO, DMAc and MDI were then sealed and kept in a suitable place until they
were required for the next synthesis.

3.3.3 Chemistry of Synthesis of Neat TPU-70 HS

The synthesis process of TPUs is susceptible, especially to humidity. The presence of


even a small amount of moisture could lead to failure of the resultant polymer or
produce a polymer with undesirable properties. Therefore, the process of
synthesising TPU met with numerous problems during this study, related to water
contamination, changing the supplier to one of the chemical compositions for
reactant materials, or stopping supplying polyol or DMAc.

Generally, in TPU synthesis, the exothermic reaction between di-isocyanate (N=C=O


groups) and the polyol groups (OH-groups) is the base of the synthesis process of
linear thermoplastic polyurethanes. In the reaction described here, there were no sub-
products with low molecular weight so that this polymerisation could be classified as
a step-wise addition polymerisation. The route used for TPU synthesis was the pre-
polymer route. The reaction was performed in two steps, which can be summarised
as follows:

In the first step, the 4,4-MDI was melted in a vessel at 80 C in an oil bath and the
polyol was added as drops by using a pressure-equalising addition funnel. The polyol
drops were continually added for 20-30 minutes with continuous stirring in order to
attain a terminal reaction between the isocyanate group and the hydroxyl group.
Once the drops of polyol had been added, the mixture was left under N2 and stirred
for 2 hours in order to ensure that the reaction was complete and that a pre-polymer
was produced. After finishing step one, the resultant product (pre-polymer) was
poured into a glass jar and sealed, then kept in a preheated oven at 50 C. In the
second step, the chain extender, 1,5 PD, which was used to extend the pre-polymer
chains, was put in the vessel with the DABCO (catalyst 1,4-
diazabicyclo[2.2.2]octane) and the solvent (DMAc) in an oil bath at 80 C.
Meanwhile, a sufficient amount of pre-polymer was mixed with an additional
amount of 4,4’-MDI and solvent inside the glass jar at 80C. Afterwards, the mixture
was poured into pressure-equalising addition funnels in order to add the mixture to
the reactant vessel in drop form. The dropping process was conducted for one hour,
and then the reactant materials were stirred vigorously under N2 gas for two hours.

85
Finally, the resultant TPU was taken and kept in sealed glass away from light until it
was required for casting and moulding. The two steps chemical reactions of TPU-70
HS synthesis are illustrated in Figure 3.6. The synthesis process set up during first
and second polymerisation process of TPU is pictured in Figure 3.7.

A) Step one: Synthesis of pre-polymer

B) Step two: Synthesis of TPU

Figure 3.6: Chemical reaction of TPU synthesis. A) Step one: synthesis of pre-
polymer and B) step two: synthesis of TPU-70 HS

86
Motor

Water in

Additional Reactant

Condenser

Water out

N2, inlet

Thermocouple

Figure 3.7: Synthesis process set-up during first and second steps of TPU-70 HS
polymerisation process

3.3.4 Calculations for Formulation of TPU

To summarise how the amounts and molarity for the TPU constituents were
calculated, the following equation can be taken as a starting point:

𝒎𝑻𝑷𝑼 = 𝒎𝑯𝑺 + 𝒎𝑺𝑺 ......... (3.1)

Where: 𝑚 𝑇𝑃𝑈 = total mass of TPU

𝑚𝐻𝑆 = mass of hard segment

𝑚𝑆𝑆 = mass of soft segment which represents the mass of polyol, so that:

𝑚𝑆𝑆 = 𝑚𝑃𝑜𝑙𝑦𝑜𝑙

However, the mass of HS represents the utility amount for both MDI in the first and
second steps (𝑚𝑀𝐷𝐼−1 ,𝑚𝑀𝐷𝐼−2 ) respectively, plus the chain extender amount

(𝑚𝐶𝐸 ), as shown in the equation below.

𝒎𝑯𝑺 = 𝒎𝑴𝑫𝑰−𝟏 + 𝒎𝑴𝑫𝑰−𝟐 + 𝒎𝑪𝑬 ......... (3.2)

87
By replacing the 𝑚𝑆𝑆 with 𝑚𝑃𝑜𝑙𝑦𝑜𝑙 and substituting the mass of HS in equation

(3.2), (3.1) for its magnitude in equation (3.1), the final mass of TPU becomes:

𝒎𝑻𝑷𝑼 = 𝒎𝑴𝑫𝑰−𝟏 + 𝒎𝑴𝑫𝑰−𝟐 + 𝒎𝑪𝑬 + 𝒎𝒑𝒐𝒍𝒚𝒐𝒍 ......... (3.3)

On the other hand, the magnitude of the pre-polymer amount which was obtained in
the first step of TPU synthesis can be calculated from:

𝒎𝒑𝒓𝒆−𝒑𝒐𝒍𝒚𝒎𝒆𝒓 = 𝒎𝑴𝑫𝑰−𝟏 + 𝒎𝒑𝒐𝒍𝒚𝒐𝒍 ......... (3.4)

As such, by substituting the magnitude of the pre-polymer mass in equation (3.4) for
its counterpart in equation (3.3), the latter can be re-written for use in the second step
of TPU synthesis, as shown below:
𝒎𝑻𝑷𝑼 = 𝒎𝒑𝒓𝒆−𝒑𝒐𝒍𝒚𝒎𝒆𝒓 + 𝒎𝑴𝑫𝑰−𝟐 + 𝒎𝑪𝑬 ......... (3.5)

In this project, the pre-polymer formulation used was estimated according to the
molar ratio (about 6:1), referring to the number of moles of MDI and polyol
respectively. There were two reasons for considering this ratio: firstly, to ensure that
there were enough isocyanate groups in the MDI to react with the hydroxyl groups in
the polyol, so that other active cyanate groups were left in the pre-polymers in order
to continue building the HS structure. The second reason was to allow the remaining
isocyanate group to react with the CE and excess MDI in step 2.

Isocyanate index (ISOindex)

For the synthesis of TPU, an understanding of the ISO index is crucial. The ISO
index refers to the amount of isocyanate required to react with the hydroxyl groups in
the polyol, calculated in terms of their stoichiometric equivalents. Therefore, the ISO
index can be calculated as follows:
𝒏𝑵𝑪𝑶
𝑰𝑺𝑶𝑰𝒏𝒅𝒆𝒙 % = ......... (3.6)
𝒏𝑶𝑯

Where: 𝑛𝑁𝐶𝑂 = number of moles of NCO

𝑛𝑂𝐻 = number of moles of OH units

This equation assumes that all the isocyanate groups react with the OH group so that
the ISO index equals one, but in this project the ISO index equals = 1.02, indicating
that there are excessive amounts of MDI at the end of the reactions in order to
continue the reaction in the second step.

88
𝒏𝑵𝑪𝑶
𝑰𝑺𝑶𝑰𝒏𝒅𝒆𝒙 % = = 𝟏. 𝟎𝟐 ........ (3.7)
𝒏𝑶𝑯

Hence the number of moles of OH groups during TPU synthesis refers to the number
of moles of OH groups in both the polyol and the chain extender, as shown by the
following equation:

𝒏𝑶𝑯 = 𝒏𝑶𝑯−𝒑𝒐𝒍𝒚𝒐𝒍 + 𝒏𝑶𝑯−𝑪𝑬 ......... (3.8)

By substituting the magnitude of 𝑛𝑂𝐻 in equation (3.8) into equation (3.7), the latter
becomes:
𝒏𝑵𝑪𝑶
𝑰𝑺𝑶𝑰𝒏𝒅𝒆𝒙 % = 𝒏 ......... (3.9)
𝑶𝑯−𝒑𝒐𝒍𝒚𝒐𝒍 +𝒏𝑶𝑯−𝑪𝑬

From equation (3.9), the magnitude of the number of moles of the NCO groups can
be obtained as follows:

𝒏𝑵𝑪𝑶 = 𝑰𝑺𝑶𝑰𝒏𝒅𝒆𝒙 × (𝒏𝑶𝑯−𝒑𝒐𝒍𝒚𝒐𝒍 + 𝒏𝑶𝑯−𝑪𝑬 ) ......... (3.10)

Generally, the number of moles (n) can be estimated using the following equation:
𝒎
𝒏= 𝒇×𝑴 ......... (3.11)

Where: f = functionality of components

M = molar ratio (g/mol)

m = mass

According to the above equation, the number of moles in the OH groups in the chain
extender (diol) and the number of moles in the NCO groups in the MDI can be
calculated as shown below:
𝒎
𝒏𝑶𝑯−𝑪𝑬 = 𝒇𝑪𝑬 𝑴𝑪𝑬 ......... (3.12)
𝑪𝑬

Where: 𝑓𝐶𝐸 = functionality of chain extender

M = molar mass of chain extender

And similarly, for MDI:


𝒎
𝒏𝑵𝑪𝑶 = 𝒇𝑴𝑫𝑰 × 𝑴𝑴𝑫𝑰 ......... (3.13)
𝑴𝑫𝑰

Where: 𝑓𝑀𝐷𝐼 = functionality of MDI

𝑀𝑀𝐷𝐼 = molar ratio of MDI

89
The magnitude of the number of moles of the hydroxyl group in one gram of polyol
is estimated according to its hydroxyl value (OHv). The OHv can be defined as the
number of milligrams of potassium hydroxide (KOH) required to titrate 1 gram of
the substance (in this case, polyol).

Therefore, the number of moles of OH per 1 gram of polyol can be written as:
𝒎
𝒏𝑶𝑯−𝒑𝒐𝒍𝒚𝒐𝒍 = 𝑴𝑲𝑶𝑯 ......... (3.14)
𝑲𝑶𝑯

Where: 𝑚𝐾𝑂𝐻 = mass of KOH

𝑀𝐾𝑂𝐻 = molecular weight of KOH in milligrams = 56100 mg/mole

But the OHv is expressed as 𝑚𝐾𝑂𝐻 in mg per 1 gram of polyol, so that equation
(3.14)

can be derived as:


𝑶𝑯𝒗×𝒎𝒑𝒐𝒍𝒚𝒐𝒍
𝒏𝑶𝑯−𝒑𝒐𝒍𝒚𝒐𝒍 = .......... (3.15)
𝑴𝑲𝑶𝑯

By substituting equations (3.15), (3.12) and (3.13) for their equivalent values in
(3.10), the resultant equation becomes:
𝒎𝑴𝑫𝑰−𝑻 𝒎 𝑶𝑯𝒗 ×𝒎𝒑𝒐𝒍𝒚𝒐𝒍
𝒇𝑴𝑫𝑰 × = 𝑰𝑺𝑶𝑰𝒏𝒅𝒆𝒙 × [𝒇𝑪𝑬 × 𝑴𝑪𝑬 + (𝑴 )] ......... (3.16)
𝑴𝑴𝑫𝑰 𝑪𝑬 𝑲𝑶𝑯 𝒊𝒏 𝒎𝒊𝒍𝒍𝒊𝒈𝒓𝒂𝒎𝒔

The magnitude of the total mass of MDI (mMDI-T)can be calculated by replacing the
mass of the chain extender with its relation with MDI.

𝒎𝑪𝑬 = 𝒎𝑯𝑺 − 𝒎𝑴𝑫𝑰−𝑻 ......... (3.17)


However,

𝒎𝑴𝑫𝑰−𝑻 = 𝒎𝑴𝑫𝑰−𝟏 + 𝒎𝑴𝑫𝑰−𝟐 ......... (3.18)


As such, the total magnitude of MDI amounts can be calculated from the amount of
MDI in steps one and two; therefore, the mass of MDI can be estimated using the
following equation:
𝒎
𝒎𝑴𝑫𝑰−𝟏 = 𝒎𝑺𝑺 × [ 𝒎 𝑴𝑫𝑰(𝒑𝒓𝒆−𝒑𝒐𝒍𝒚𝒎𝒆𝒓 𝒇𝒐𝒓𝒎𝒖𝒍𝒂𝒕𝒊𝒐𝒏) ] ......... (3.19)
𝒑𝒐𝒍𝒚𝒐𝒍 (𝒑𝒓𝒆 𝒑𝒐𝒍𝒚𝒎𝒆𝒓 𝒇𝒐𝒓𝒎𝒖𝒍𝒂𝒕𝒊𝒐𝒏)

The amounts of both mMDI (pre-polymer formulation) and 𝑚𝑝𝑜𝑙𝑦𝑜𝑙 (𝑝𝑟𝑒 𝑝𝑜𝑙𝑦𝑚𝑒𝑟 𝑓𝑜𝑟𝑚𝑢𝑙𝑎𝑡𝑖𝑜𝑛) ) are
estimated from the molar ratio 6:1 for the pre-polymer formulation for both the MDI
and the polyol respectively.

The mass of MDI in the second step can then be calculated as shown below:

90
𝒎𝑴𝑫𝑰−𝟐 = 𝒎𝑯𝑺 − 𝒎𝑴𝑫𝑰−𝟏 − 𝒎𝑪𝑬 ........ (3.20)

Numerical calculation of TPU-70% HS synthesis can be used as an applied


example for the above TPU formulation calculation.

To synthesise 100g of TPU-70% HS, the calculation of the amounts of its chemical
reactant materials can be formulated as shown below:

The amount of 𝑚𝑀𝐷𝐼−𝑇 required for the synthesis of TPU-70% HS can be


calculated using equation (3.16):
mMDI−T CE m 56.1 ×30
2× = 1.02 × [2 × 104.15 +( )]
250.25 56100

2 2.04 56.1×30×1.04
× 𝑚𝑀𝐷𝐼−𝑇 = × (𝑚𝐻𝑆 − 𝑚𝑀𝐷𝐼−𝑇 ) +
250.25 104.15 56100

2 2.04 2.04×70 56.1×30×1.02


× 𝑚𝑀𝐷𝐼−𝑇 + 104.15 × 𝑚𝑀𝐷𝐼−𝑇 = +
250.25 104.15 56100

1.40170
𝑚𝑀𝐷𝐼−𝑇 = 0.02758

𝑚𝑀𝐷𝐼−𝑇 = 50.8246 𝑔

The amount of chain extender can be calculated from equation (3.17) using the total
mass of MDI:

𝑚𝐶𝐸 = 70 − 50.8246

𝑚𝐶𝐸 = 19.16 𝑔

Step 1- Pre-polymer synthesis

𝒎𝑴𝑫𝑰−𝟏 can be calculated using equation (3.19)

when:

64.29
𝑚𝑀𝐷𝐼−1 = 30 × [ ]
85.634

𝑚𝑀𝐷𝐼−1 = 22.522 g

Step 2

From equation (3.20), the amount of 𝑚𝑀𝐷𝐼−2 in step 2 can be calculated as


follows:

91
𝑚𝑀𝐷𝐼−2 = 70 − 22.522 − 19.16

𝑚𝑀𝐷𝐼−2 = 28.318 𝑔

To calculate the total amount of 𝑚 𝑇𝑃𝑈 , equation (3.4) can be used:

𝑚 𝑇𝑃𝑈 = 22.522 + 28.318 + 19.16 + 30 = 100 g

3.4 Synthesis of TPU70-NCs using In situ Polymerisation

In the present work, the TPU70-NCs were prepared using in situ polymerisation.
This method is widely used to improve interfacial interaction between polymer and
fillers and to achieve good filler dispersion. It has been demonstrated previously in
the literature review that the in situ polymerisation technique gives better dispersion
of nanofiller within a polymer matrix in comparison with other methods [46, 48, 152-
154]. Therefore, the synthesis of TPU70-NCs for all three types of GNPs was
achieved using in situ polymerisation. This process depended on adding nanofillers
during the polymerisation process. In this case, amounts of each GNPs types (M5,
M15 and M25) were mixed with the solvent (DMAc) and dispersed using an ultra-
sonicated water bath at a frequency of 37 kHz and a temperature of 30 C for
approximately two hours. From Figure 3.8, the disparity can be observed between the
dispersion of GNPs within DMAc before and after the ultra-sonicated water bath
process, which confirms the effect of sonication on dispersing the GNPs particles
within the solvent.

(A) (B)
Figure 3.8: Dispersion of GNPs in DMAc solvent using an ultra-sonicated water
bath technique: A) before sonication and B) after sonication

92
Then the homogeneous dispersion solution (DMAc+GNPs) was directly added to the
chain extender inside the vessel during the second stage of the TPU-70 HS
polymerisation process in order to prevent re-stacking of the GNPs flakes and
subsequent agglomeration. Later, drops of the mixture of MDI, pre-polymer, and
DMAc were added in order to stimulate the reaction between the GNPs and the chain
extender during polymerisation. Once the second step was complete, the resultant
nanocomposite solution was poured directly into PTFE moulds and dried in the oven
in order to prevent separation of the nanoparticle from the TPU solvent and to obtain
a casting film of TPU70-NCs solution.

3.4.1 Solvent-casting Films

The moulds used for casting the TPU-70 HS and TPU70-NCs solution were made
from Teflon (polytetrafluoroethylene (PTFE)) with a thickness of 20 mm. The frame
that was made inside the mould had a cross-section area measuring 75 mm  250 mm
and a thickness of 10 mm on both sides. In order to dry the TPU-70 HS and TPU70-
NCs, a sufficient amount of them was poured into the casting mould and placed
inside a pre-heated oven at 80 C for 3 days in atmospheric air. On the fourth day,
the samples were dried under a vacuum for six hours to ensure that all the solvent
had evaporated. Finally, the films of TPU-70 HS or TPU70-NCs were obtained with
a thickness of around 1-2 mm. These films were then taken out, kept inside a plastic
bag, and stored in the desiccator for later use. Figure 3.9 shows the drying films of
TPU-70 HS and TPU70-NCs after removal from the oven.

(A) (B)

Figure 3.9: Solvent-casting of A) TPU-70 HS and B) TPU70-NCs films in PTFE


moulds

93
3.5 Manufacturing Process

3.5.1 Injection-moulding (IM) Process

Injection moulding was one of the manufacturing processes used in this work to
prepare TPU-70 HS and TPU70-NCs samples for tensile testing. The instrument
used to achieve this task was the Haake Minijet II (Thermo-Scientific), operating at a
barrel temperature of 200 °C, a mould temperature of 60 °C, an injection pressure of
850-1100 bar for 10 seconds and a holding pressure of 400 bar for 5 seconds. The
resultant moulded specimens were dumbbell-shaped with dimensions according to
the ISO 527-2 type 1BA standard (30 mm gauge length, and width and thickness of
approximately 4.5 mm and 1.5 mm respectively) [155]. Figure 3.10 shows the
Haake Minijet injection-moulding machine which was used in the present work, as
well as the dumbbell-shaped mould and some of the resultant nanocomposite tensile-
test samples.

(B) (C)
(A)
Figure 3.10: A) Haake Minijet II injection-moulding machine, B) dumbbell-shaped
mould, and C) some of the resultant nanocomposite samples

3.5.2 Compression-moulding (CM) Process

Compression-moulding was another manufacturing process used to prepare neat


TPU-70 HS and TPU70-NCs samples for thermal and electrical conductivity testing.
This process was also used to fabricate the panel from TPU70/CF laminates and

94
TPU70/CF-multiscale composites (MSCs) in order to conduct mechanical, electrical
and thermal testing, as illustrated below.

3.5.2.1 Compression-moulding for Preparing Samples of Neat TPU-70


HS and TPU70-NCs
The neat TPU-70 HS and TPU70-NCs moulded samples for measuring rheological
and thermal behaviour, thermal and electrical conductivity were made using
compression-moulding. The compression-moulding process was carried out in two
stages that can be summarised as follows:

First stage of compression-moulding process

In this stage, the solvent-casting film for both materials was converted to a thin film
by flatting them at their melting temperature (174 °C for unfilled TPU-70 HS and
182-185 °C for the TPU70-NCs) using a hydraulic hot press machine at pressures
between 25-35 bar (around 2.5-3.5 MPa). This procedure was performed by placing
the neat TPU-70 HS film or its nanocomposite between two PTFE-released films in
order to prevent them from sticking inside the hot press plates. After 2-3 minutes the
hot press was opened and the film was taken out and left to cool at room temperature.
Typically, the resultant flattened film was approximately 0.2-0.3 mm thick,
depending directly on the melting temperature and pressure that was used.

Second stage of compression-moulding process

This stage included pre-heating the mould to the melting temperature of neat TPU-70
HS and the nanocomposite; during this time, the flattened film was cut into pieces
according to mould size. For example, the dimensions of the mould used to prepare
samples for the thermal and electrical conductivity tests were 70 mm  70 mm  4
mm. After around fifteen minutes, the heated mould was taken out, the flattened
TPU-70 HS or nanocomposite films were placed in it, and it was placed inside the
heated hydraulic hot press machine. The pressure that was used to compress the
sample was around 10-14 MPa, depending on the mould area. The samples of neat
TPU-70 HS and nanocomposite were kept at this pressure for around 5-10 minutes,
and then the mould was cooled to 25 °C using a water-cooling system at a cooling
rate of approximately 20 °C/min, as shown in Figure 3.11. Finally, the samples were
taken out of the mould and immediately placed in plastic bags in order to minimise
water uptake.

95
200 160

180 140
160
120

Temperature (C)

Pressure ( bars)
140
100
120
80
100
80 60

60 40
40
20
20
0
0 200 400 600 800
Time (Sec)

Figure 3.11: Schematic representation of the cyclic hot-press consolidation process


used for manufacturing neat TPU-70 HS and TPU70NCs samples

3.5.2.2 Preparation of Multiscale Composites (MSCs)


The TPU70/CF laminate and MSCs were prepared using compression-moulding in
the same way in which the NCs and neat TPU-70 HS samples were prepared, except
for differences in the pressure and time used for compression and the arrangement of
the CF and matrix layers. It is essential to find the suitable temperature, time and
pressure in order to prevent defects, such as voids, bubbles, and the like, from
forming inside the resultant composite. In general, this process includes two stages;
the first stage also involves flattening the neat TPU-70 HS and NCs, as mentioned
previously (see section 3.5.2.1). The second stage included the manufacturing of
TPU70/CF laminate and MSCs. Before beginning the compression-moulding
process, the carbon fibres were cut by sticking all of their edges using masking tape
in order to keep the yarns facing in the same direction after cutting, without any
deterioration being caused by the size of the mould. The mould was then pre-heated
to the matrices’ melting temperatures. The CF, neat TPU-70 HS and NC matrix films
were cut according to the size of the mould, which varies from one test to another
depending on the ASTM standards deemed appropriate. The pre-heated mould was
then taken out and the CF and layers of the matrices were alternately laid out by hand
within it (as illustrated in Figure 3.12). For instance, for 2 mm thickness, 4 CF layers
and 5 matrix layers were used. After that, the mould was returned to the hot-press
machine and a pressure of about 12-14 MPa was applied, depending on the mould
area, for 5-10 minutes. In fact, the compressed pressure that was used in this study to

96
produce testing samples of TPU70/CF and MSCs followed the pressure that was
mentioned and used by Al-Obad [47] study to produce TPU70/glass fibre
composites. Finally, the process of cooling the composite took place in the same way
mentioned in section 3.5.2.1. Finally, the resultant composites were preserved in a
plastic bag to be cut later.

Steel CF layer Resultant


Steel
mould composite
plates Matrix layer

(A) (B) (C)


Figure 3.12: Schematic representation of the compression-moulding process for
TPU70/CF laminate and multiscale composite: A) preheated mould, B) arrangement
of TPU films and CF layers, C) resultant composite

3.6 Cutting Process for Preparing TPU70/CF Laminate


and MSCs Samples

After fabricating the TPU70/CF laminate and MSCs panels, the specimens were cut
according to the dimensions of the standards used for thermal and mechanical tests.
The cutting process was carried out using a water-cooled diamond-cutting wheel for
all tests except for the SEM specimens, which were cut using a precision cutting
machine. This tool was used in order to attain a smoother surface compared with the
previous cutter. After cutting, the specimens were left to dry out at room temperature
and then placed in an oven overnight at 50 °C. On the next day, the specimens were
taken out and kept inside a plastic bag within a desiccator until they were required
for use in the tests.

3.7 Annealing Treatment

To prepare annealed samples for thermal and mechanical testing, the same numbers
for each test of neat TPU-70 HS, TPU70-NCs, TPU70/CF and TPU70-MSCs
samples before and after annealing were selected after preparation and cutting
process. The annealing treatment was carried out by putting the samples between two
steel plates in order to avoid bending. Then, the two plates and samples were put in
the oven at 80 C for three days. Finally, the plates were taken out and the samples

97
were left to cool to room temperature then kept in the plastic bags within a desiccator
until they were used for testing.

3.8 Characterisation and Experimental Procedure

3.8.1 Gel Permeation Chromatography (GPC)

Background of GPC

The physical and mechanical properties of polymers are determined by their


molecular weight [70, 156]. With an increase in molecular weight, some properties
increase, such as melting temperature, viscosity and mechanical strength [70, 157].
Conversely, other properties collapse; for instance, elongation at break and melt flow
index [158]. Therefore, for structural applications, the polymer should have robust
mechanical properties, and this is only reachable when their molecular weight is
greater than 10,000 g/mol [156]. Otherwise, a low molecular weight or thin film is
sometimes required for special applications. Figure 3.13 illustrates how the
mechanical strength of a polymer is considerably affected with an increase in
molecular weight.

Figure 3.13: Influence of molecular weight on the mechanical strength of a polymer


[156, 157]
Point A in this figure represents the minimum molecular weight of a polymer, which
is usually around a thousand. After point A, the mechanical strength starts to increase
swiftly with increasing molecular weight until it reaches a critical point (B). Point B
represents the minimum molecular weight that is required to achieve a polymer with
adequate mechanical strength. The minimum molecular weight varies from one
polymer to another, but generally falls within the range of 5,000 to 10,000 g/mol [70,

98
156, 157]. Above point B, the mechanical strength increases gradually and finally
reaches point C, which represents a limiting value.The number-average molecular
̅𝑛 can be calculated using equation (3.21) [70, 156, 157]:
weight 𝑀

̅ 𝒏 = ∑ 𝑵𝒙𝑴𝒙
𝑴 ......... (3.21)
∑ 𝑵𝒙

where 𝑁𝑥 represents a number of molecules whose molecular weight is 𝑀𝑥 , and x


represents a number from 1 to . Likewise, the weight-average molecular weight
̅𝑤 can be calculated by equation (3.22):
𝑀
𝟐
̅ 𝒘 = ∑ 𝑵𝒙𝑴𝒙
𝑴 ......... (3.22)
∑ 𝑵𝑿 𝑴𝒙

̅𝑛 is always lower than the 𝑀


The value of 𝑀 ̅𝑤 value, except in the monodisperse
system when their values are equal [157, 159]. The molecular weight polydispersity,
or PDI, as it is known, is calculated according to the following ratio:

̅𝒘
𝑴
𝑷𝑫𝑰 = ̅𝒏
......... (3.23)
𝑴

The polydispersity represents a measurement of the extent of polymer molecular


weight distribution. Therefore, the gel permeation chromatography (GPC) or size
exclusion chromatography (SEC) technique are considered the best methods with
̅𝑛 and 𝑀
which to measure 𝑀 ̅𝑤 and calculate PDI using only one sample [156, 159].
GPC can be defined as the separation mechanism that depends on the size of the
polymer molecules in the solution and determines the distribution of the average
molecular weight of polymers [70].

GPC Experimental Procedure

GPC is a liquid chromatographic technique that depends on dissolving the polymer


in an appropriate organic solvent and separating the solution based on the size of the
molecules. In this technique, the solution is first injected into the column that is
packed with porous particles [70, 159]. An organic solvent (mobile phase) is used to
carry the polymer chains through the column. The separation technique starts when
the large molecule chains are eluted first due to them being excluded from the pores.
Conversely, the smaller polymer chains can create a longer flow path by moving into
most of the pores; consequently, a higher retention time is obtained, as illustrated in
Figure 3.14.

99
Figure 3.14: Separation process in GPC column based on disparity in molecular size
[70, 159]
In this work, the preparation of GPC samples of TPU-70 HS was performed as
follows: around 10 mg of TPU polymer was dissolved in 10 ml of tetrahydrofuran
(THF-Sigma-Aldrich) with a ratio of 1 mg/ml. The mixture was then placed in a
magnetic glass jar and stirred around 5-7 hours at 30 °C. The solution was then
filtered into blue-capped HPLC tubes using a syringe filter (Millex – HV, Durapore,
made from 0.45-micron PVDF). A Shodex RI 101 refractive index detector was used
along with the GPC instrument. In the stationary phase, phenogel columns with a
polymer porous beads or particles (such as poly (styrene-co-divinylbenzene)
particles) size of 5 microns were used. During measurement, the flow of the solutions
was maintained at 1 ml/min and around 0.5 ml of solution was injected into the
column.

3.8.2 Differential Scanning Calorimetry (DSC)

Background of DSC

DSC is a thermal analysis technique used with materials that are subjected to a
chemical reaction and/or phase transition during heating or cooling [52]. This
process was accomplished by recording the change in heat flow between the
reference and the sample with respect to temperature or time during heating and

100
cooling. The results permit observation of aspects of thermal transition, such as
melting temperature (Tm) (when the temperature increases), crystallisation
temperature (TC) (when temperature decreases) and glass transition temperature (Tg).
Besides, the results give an indication of the enthalpy, curing and heat capacity of the
material [160, 161].

There are two types of DSC techniques: heat flux and power compensation DSC. In
the heat flux technique, which is also known as ‘quantitative DTA’, the difference
between the reference and sample temperatures is directly measured and then
converted to the change in heat flow (enthalpy change), as shown in Figure 3.15a.
The conversion is achieved via an algorithm in the computer software installed in the
system [162]. In power compensation DSC, each sample and reference are placed in
two separate heating chambers where their temperatures are controlled independently
(Figure 3.15b). When a thermal event occurs in the sample materials, power will
change in one or both of the chambers in order to compensate for this change and
maintain the sample temperature [161, 162].

This project adopted the heat flux technique, which is more common than the power
compensation technique [161]. In both types, a metal pan is used to hold the samples
and reference, where the pan used for the latter is normally empty or can contain an
inert material. However, to obtain reliable results, it is recommended that the same
metal pan is used for the samples and reference. Typically, pans made from
aluminium are utilised in this technique, which can be used at temperatures ranging
from -150 to 600 C [161].

(a)

Figure 3.15: Design of DSC instruments: a) heat flux DSC, and b) power
compensation DSC; A) furnace, B) separate heater and C) samples and reference
holders [162]

101
The heating and/or cooling of the sample and reference can be done at a constant
𝑑𝑞
rate, and the heat flow can be written as , where q represents heat and t indicates
𝑑𝑡
time. When the DSC experiments are run at a constant pressure, the heat flow can be
written as:

𝒅𝒒 𝒅𝑯
( 𝒅𝒕 ) = 𝒅𝒕
......... (3.24)
𝒑

𝑑𝐻
where represents the change in enthalpy over time, measured in Joule/s.
𝑑𝑡

Therefore, in DSC experiments the change in heat flow between a sample and a
reference can be written as:

𝒅𝑯 𝒅𝑯 𝒅𝑯
∆ = ( ) − ( 𝒅𝒕 ) ......... (3.25)
𝒅𝒕 𝒅𝒕 𝒔𝒂𝒎𝒑𝒍𝒆 𝒓𝒆𝒇𝒆𝒓𝒆𝒏𝒄𝒆

The peaks that can be observed in the outcome plot from the DSC experimental
results represent exothermic transition, such as crystallisation (heat flow decrease),
and endothermic transition, such as melting (heat flow increase). The total change in
enthalpy (H) can be calculated from the area under these peaks:

𝒅𝑯
∫ ( 𝒅𝒕 ) 𝒅𝒕 = ∆𝑯𝒔𝒂𝒎𝒑𝒍𝒆 ......... (3.26)
𝒔𝒂𝒎𝒑𝒍𝒆

During thermal transitions such as glass transition, the change in the heat capacity
can also be observed as the steps in the power signal baseline; these steps are
discontinuities [52, 163]. Thus, the heat capacity under constant pressure can be
written as:

𝒅𝒒 𝒅𝑯
𝑪𝒑 = ( ) = ( ) ......... (3.27)
𝒅𝑻 𝒅𝑻 𝒑 𝒑

where 𝐶𝑝 is the heat capacity measured in J/g .°C

For further analysis, the heat capacity can be rewritten depending on the heat scan
𝑑𝑇
rate ( ):
𝑑𝑡

𝒅𝑯 𝒅𝑯 𝒅𝒕
𝑪𝒑 = ( ) = ( 𝒅𝒕 ) (𝒅𝑻) ......... (3.28)
𝒅𝑻

From equation (3.27):

102
∆𝑯 = ∫ 𝑪𝒑 𝒅𝑻 ......... (3.29)

DSC Experimental Procedure

The DSC experiment was carried out using a TA Q100 instrument supplied with an
auto-sampler and used the Indium calibration standard. Before commencing the test,
all of the empty pans and samples of TPU-70 HS, TPU70-NCs, and TPU70/CF
laminate and TPU70/CF-MSCs composite were weighed before and after annealing
in order to record any change in weight during the test as a result of being subjected
to the subsequent thermal protocol. The weight of the samples was around (7-11 ±
0.5mg) and at least three samples were used for each material. After weighing, all of
the samples were sealed in hermetic aluminium pans and the investigation was
carried out under an N2 atmosphere. The heating protocol for a custom type based on
a cool/heat/cool/heat sequence with a heating range of -90 to 220 C and rate of
10C/min is summarised in Table 3.8. The TA Universal Analysis software was used
to analyse all of the DSC, DMTA and TGA results.

Table 3.8: Conventional DSC experiments’ thermal protocol for non-annealed and
annealed TPU-70HS, TPU70/GNPs-NCs, TPU70/CF laminate and TPU70/CF-
MSCs.

Cycle Sequence Isotherm time


Cycle 1 Cooling from 25 to -90 C 5 min at - 90 C
Heating from -90 to 220 C at rate of 10
Cycle 2 2 min at 220 C
C/min
Cycle 3 Cooling from 220 to -90 C 5 min at -90 C
Heating from -90 to 220 C at rate of 10
Cycle 2 min at 220 C
C/min
The results obtained during the first heating cycle represent the previous thermal
history of all of the samples through their casting process, while a second heating
cycle was applied to investigate the thermal changes in the samples.

3.8.3 Dynamic Mechanical Thermal Analysis (DMTA)

Background of Dynamic Mechanical Thermal Analysis (DMTA)

DMTA is a technique used to determine the viscoelastic properties of various


materials including polymers. It is appropriate for measuring dissipation energy,
dynamic moduli, mechanical damping and glass transition [70, 164]. In DMTA, the
sample is subjected to sinusoidal oscillating stress which induces the corresponding

103
sinusoidal strain and phase difference (damping - ). The phase difference occurs
when the resultant strain lags behind the applied stress, as illustrated in Figure 3.16.
For example, the sample is utility-elastic if the phase difference is equal to zero,
indicating that both stress and strain are in the same phase; alternatively, if  = 90,
this means that the sample is purely viscous and the stress leads the strain to 90° out
of the phase. On the other hand, if the  lies between 0° and 90°, this indicates a
sample with viscoelastic behaviour [164-166].

Figure 3.16: The relation between stress and strain during dynamic mechanical
testing [70, 165, 166]
The dynamic stress and strain are calculated according to the following equations
[166]:

(𝒕) = 𝒎𝒂𝒙 𝒔𝒊𝒏 ( + ) ......... (3.30)

(𝒕) = 𝒎𝒂𝒙 𝒔𝒊𝒏  ......... (3.31)

where: (t) and (t) = stress and strain at time t respectively

𝑚𝑎𝑥 and 𝑚𝑎𝑥 = maximum stress and strain


 = angular frequency = 2 f, and f = frequency in Hz

 = phase difference

The peak stress-to-peak strain ratio represents the complex modulus (*), which has
two components: the storage modulus or in-phase component (’) and the loss
modulus (”) or out-of-phase component. The storage modulus is derived from

104
elastic material behaviour and relates to the recoverable energy, while loss modulus
refers to the viscous material behaviour and relates to the dissipation of energy due to
deformation. The complex modulus and its components can be represented in a
diagram known as the Argand diagram, shown in Figure 3.17 [70, 164, 166].

Figure 3.17: Argand diagram to explain the relation between complex modulus and
its components [70, 165, 166]
From this diagram, the magnitude of the loss modulus and storage modulus can be
derived as:

 ′ = ∗ 𝒄𝒐𝒔 ......... (3.32)

 = ∗ 𝒔𝒊𝒏  ......... (3.33)


In order to represent a vector quantity, the complex modulus has a real vector
component () and an imaginary component (), identified by , as shown in the
following equations [164, 166]:
(𝒕)
∗ = (𝒕)
=  + 𝒊 ......... (3.34)

Furthermore, tan  represents the ratio of loss modulus to storage moduli (equation
(3.35)). This ratio is known as the mechanical damping factor due to its sensitivity to
molecular mobility, which in turn influences the mechanical properties of materials
[167].


𝒕𝒂𝒏() = ......... (3.35)


DMTA Experimental Procedure

All of the experiments to investigate the thermo-mechanical properties of neat TPU-


70 HS, TPU70-NCs, TPU70/CF laminate and MSCs before and after annealing were

105
carried out using a DMA Q800 from TA Instruments. This instrument used liquid
nitrogen as a gas-cooling accessory (GCA). The test was carried out using the Dual
Cantilever clamping mode in order to examine the neat TPU-70 HS and its
nanocomposites and the multiscale composite. The dimensions of the Dual
Cantilever specimen were roughly 60 mm in length, 12.7 mm wide and 3 mm thick.
The conditions for this test consisted of a temperature ramp from -120 °C to 150 °C
with a heating rate of 3 °C/min and amplitude of 10 m. Measurements were
performed by applying an oscillatory stress at 1Hz constant frequency using three
samples for each type of different material. The outcome data was displayed as
dynamic storage modulus (E’) and damping factor (tan) curves. The storage
modulus can give information about the matrix reinforcement owing to the existence
of inorganic filler, while from the highest peak on tan  curve; the Tg of polymer
nanocomposite can be estimated, giving an indication to the effect of nanofillers on
the mobility of the polymer macromolecular chain.

3.8.4 Thermogravimetric Analysis (TGA)

Background of TGA

TGA is considered one of the oldest and simplest techniques for thermal description.
This technique measures the weight change of the sample as a function of
temperature when the sample is subjected to a constant rate of heat under a
controlled atmosphere. The atmosphere could be reactive, such as air and oxygen, or
non-reactive (inert gas), such as N2 and argon [168]. Primarily, it is used to study the
thermal stability and decomposition of materials, as well as to observe the existence
of any volatile components. Typically, the results of TGA are presented as a plot of
weight loss on the y-axis towards temperature or time on the x-axis; the shape of this
plot is sigmoidal in nature [169]. Despite most weight loss occurring at the main
reaction temperature, there is still residual weight loss afterwards, as shown in Figure
3.18A. Alternatively, the results of TGA can be represented as another plot when the
rate of weight loss against temperature (dm/dT) is plotted on the y-axis and
temperature remains on the x-axis. In this case, the test is called Derivative
Thermogravimetric Analysis (DGA), as shown in Figure 3.18 B, and the plot shows
a high peak above a range of temperatures. This plot is highly useful during
overlapping reactions, as it offers shoulders or a double peak [169, 170].

106
(A) (B)
Figure 3.18: A) TGA curve, and B) DGA curve [70, 169]
TGA Procedure

All TGA experiments were performed using a modulated TGA Q500 thermo-
gravimetric analyser from TA Instruments. The sample was placed in a platinum pan
and the experiments were carried out under an N2 atmosphere. The temperature range
used was 25 to 1000 °C for graphite and GNP powder, and 25-800 °C for neat
polymers and NCs, at a heating rate of 10 °C/min; the mass of each sample was
between 5-10 mg. TGA was carried out to investigate the effect of GNPs (M5, M15
and M25) on the thermal stability of the matrix, TPU-70 HS polymer when heated
from room temperature to 800 °C.

3.8.5 Thermal Conductivity

This test was carried out to assess the ability of nanofillers to improve the thermal
conductivity of polymers. Generally, GNPs are known for their outstanding thermal
conductivity that is expected to enhance the thermal properties of polymer
nanocomposites. This enhancement, however, depends on various factors, such as
orientation, quality of the dispersion of nanofillers and their loading, and further
interaction between polymers and nanofillers.

The test was performed using a Fox 50 heat flow meter from LaserComp Inc. to
measure the thermal conductivity of TPU-70 HS and TPU70/GNPs-NCs. The
samples were prepared in two different thicknesses (2 and 4 mm) ant two samples
were used for each thickness with diameter 50 mm, and their thermal conductivity
was calculated according to ASTM C518 and ISO 8301 [171]. The sample was
placed between two plates in the test stack and a gradient of the temperature was
107
generated over the thickness of the tested sample. During the test, the upper plate
remained stable, where the pneumatic drive sets the lower stack to assure reasonable
contact between the sample and transducer. The general principle of heat flow in Fox
50 meter instruments is based on a one-dimensional equation of the Fourier-Biot
Law, as follows:

𝒅𝑻
𝒒 = −𝒌𝒕𝒉 ( ) ......... (3.36)
𝒅𝒙

where q is the heat flux (W/m2) flowing through the sample, 𝑘𝑡ℎ represents the
𝑑𝑇
thermal conductivity (W/m.K), and represents the temperature gradient (K/m) on
𝑑𝑥

the isothermal flat surface of the sample.

When a flat sample is put between two flat isothermal plates that have a different
heating temperature, a homogenous one-dimensional temperature field is generated
within the specimen volume. The gradient of temperature field within the sample is
equal to the difference between the temperatures of its surfaces ΔT (ΔT = Thot surface –
Tcold surface) divided by its thickness, Δx, because in this case the average temperature
gradient (dT/dx) in equation (3.36) is equal to (–ΔT/Δx). Therefore, the thermal
resistance of the flat sample (Rsample) is equal to its thickness Δx divided by its
thermal conductivity, kth. The principle of heat flow meter measurement was based
on combining the thermal resistance of the sample and the thermal resistance of the
interfacial surface between the sample and transducer. However, the interfacial
resistance can often be higher than the bulk thermal resistance of the sample; this
indicates that the measurement of thermal conductivity should be lower than its
actual value.

In this study two different average temperatures, 20 °C and 30 °C, were used to carry
out the test. At 20 °C, the upper plate was set to 25 °C, while the lower plate was set
to 15 °C. At 30 °C, the upper plate was set to 35 °C and the lower plate was fixed at
25 °C.

3.8.6 Wide-Angle X-ray Diffraction (WAXD)

Background of WAXD

WAXD is a method that is used to determine the crystalline structure of the materials
and their degree of crystallinity. This method consists of measuring the density of

108
electrons scattering from the structure of materials, then observing the diffraction of
X-rays at scattering angles typically between 5° and 90°. From this method,
information on d-distance can be attained from 1 to 10 Å.

Diffraction occurs if the scattering electrons have satisfied Bragg’s Law when they
take place in the atomic lattice. Some polymers have a semi-crystalline structure,
which is a combination of amorphous and crystalline structures. Therefore, the
WAXD spectrum of a semi-crystalline polymer sample contains two different kinds
of peaks: “Bragg peaks” and diffuse scattering peaks. The Bragg peaks can be
observed as sharp distinctive peaks with high intensity, which results from the
crystalline structure of solids. The diffuse scattering peaks, known as an “amorphous
halo”, indicate the amorphous region [52].

In addition, in the crystalline region of a semi-crystalline polymer sample under


examination, the diffraction pattern contains a discrete reflection of Bragg peaks,
which consists of information about its atomic ordering structure. In crystalline or
semi-crystalline materials, the atoms in the ordered region have a function as
numerous scatters for electromagnetic radiation, meaning that an X-ray beam can
penetrate many layers of the crystalline region of a sample under observation, as
shown in Figure 3.19.

d-spacing

Figure 3.19: Schematic representation of the X-ray diffractions from the planes of a
crystal lattice found in an ordered system [52, 172]
Some parts of the X-ray incident are scattered by atoms due to the difference in
layers of crystalline regions that are separated by inter-atomic distance (d-spacing).
The scattering X-ray beams form two sets, known as constructive interference and
destructive interference. Destructive interference is usually neglected because it

109
results from scattering the X-rays in and out of phase in other angles of space.
Conversely, constructive interference produces diffracted peaks that are collected
and analysed. It is generated because the atoms scatter the incident X-rays in phase
with each other [52, 173]. Bragg’s Law can only be applied to constructive
interference. The following equation represents the Bragg’s Law equation, known as
Bragg’s equation, which gives information about the concept of X-ray diffractions as
well as the relation between the X-rays and the crystalline plane of ordered materials.

𝒏 = 𝟐𝒅𝒔𝒊𝒏  ......... (3.37)

Where:

 = wavelength of the X-rays

d = distance between the diffracting planes of the atomic lattice

 = the angle between the scattering planes and the incident X-ray

Experimental Procedure of WAXD

WAXD experiments were performed for GNPs (M5, M15 and M25) before and after
sonicating, along with neat TPU-70 HS and TPU70-NCs before and after annealing.
The PANalytical X’Pert Pro (XRD5) instrument was used to carry out this study,
which used a copper anode radiation source (Cu) (K = 1.5419Å); the conditions for
running the test were 40 m and 40 kV. The scan mode used was continuous and the
scan step time was 10 s with a scan size of 0.0334. The sample size for neat TPU-
70 HS and its nanocomposite and MSCs were 1 cm2 and 2 mm thickness. All of the
experiments were conducted at room temperature using a 2 range from 5 to 90 for
each sample.

3.8.7 X-ray Photoelectron Spectroscopy (XPS)

XPS represents a strong research tool for analysing the chemical elements of any
solid surface, quantifying and identifying their composition and/or speciation within
a depth of 1-10 nm. During XPS test, the X-ray photons are applied on the
investigated sample resulting a transformation of energy from the applied photon to
the electron within atoms or ions due to their interaction together. If this energy is
enough, it will lead to rejection of the electrons from those atoms or ions [174].
Therefore, the XPS can measure the kinetic energy (𝐸𝑘 ) of this emitted electrons.

110
This kinetic energy in general refers to the remaining energy of the emitted
photoelectrons that can be quantified experimentally [174, 175]. The initiated photon
energy (𝐸𝑖𝑝ℎ ) is already known; as such, the binding energy (𝐸𝑏 ) can be calculated
using the following equation:

𝑬𝒃 = 𝑬𝒊𝒑𝒉 − 𝑬𝒌 − 𝝋𝒔 ......... (3.38)

where 𝜑𝑠 represents the spectrometer work function of the instrument.

In this study, a Kratos Axis DLD Ultra instrument was used to analyse the GNPs as
received and after extraction from TPU-70 HS after being washed twice using
DMAc solvents. The samples were prepared for the first and second wash using a
centrifuge at a speed of 1500 rpm for 45 minutes. The GNPs were then filtered and
dried in an oven for 4 days. A silicon wafer with size 1  1 cm (width  length) was
used to substrate 2-3 drops from the DMAc and GNPs solution. The sample was then
left to dry overnight in a fume cupboard before the test was carried out.

A hemispherical analyser and a monochromatic AlK source (1486.62 eV) were


used. The survey spectra were measured using a pass energy of 80 eV in increments
of 0.5 eV and 100 ms dwell time per data point. For high-resolution spectra, the
measurement was carried out using a spectral range of interest (ca. ±20 eV around
core-level emission peaks) and a pass energy of 20 eV with 0.1 eV steps. The results
were analysed using CasaXPS software.

3.8.8 Elemental Analysis (CHNS Analysis)

CHNS analysis is a simple technique used to measure the quantity of carbon,


hydrogen, nitrogen and sulphur on the surface of as-received and extracted GNPs.
The experiment was performed in the School of Chemistry at the University of
Manchester, UK, using a CHNS Thermo Scientific Flash 2000 organic elemental
analyser. The results obtained from this test were compared with the XPS results
using the same sample and the same extraction technique.

3.8.9 Scanning Electron Microscopy (SEM)

Background of SEM

The SEM technique is one of the most commonly and widely used techniques that is
used to study the morphology of nano-fillers, polymers and polymer composites.

111
Images are produced by the SEM instrument using a beam of electrons in order to
scan the sample surface at a higher resolution. There are three different types of
interactions that may occur when the beam of electrons interacts with the sample
surface, which can be illustrated as follows [162, 176]:

I) Backscattered electrons (BSE) can be generated in an area which is deep into


the sample (around 1m or more) due to the electrostatic interaction
between the free electrons in the incident beam (negative charge) and the
nucleus within the sample (positive charge).
II) Secondary electrons (SE) can be knocked free and generated in the sample at
a depth of about 5-50 nm as a consequence of some of the electrons in the
incident beam interacting directly with electrons within the atoms of the
samples.
III) An X-ray spectrum can be emitted when some electrons with a weak bound
state fall into the inner shell after the secondary electrons are removed
from an inner shell.

These are the most common interactions and extensively used in images mode of
SEM, which are summarised schematically in Figure 3.20.

Figure 3.20: Schematic illustration of beam-sample interaction shows backscattered


electrons (BSE), secondary electrons (SE) and X-rays [176]
These types are chosen depending on the information to be collected from the SEM

technique. For example, the backscattered electrons (BSE) are useful for detecting
the contrast resolution between areas with different elemental compositions. The
secondary electrons (SE) can be used to observe surface contrast, while X-rays are
useful for both chemical investigation and elemental analysis. During this work the

112
SE mode was adopted to examine the morphology of GNPs fillers, neat TPU-70 HS
and TPU70/GNPs-NCs. In addition, the technique was used to investigate the
fracture surface of TPU70/CF laminate and multiscale composite after mechanical
testing.

SEM Sample Preparation and Experimental Procedure

A study of the effect of the sonication process on as-received GNPs (M5, M15 and
M25) was carried out using a Zeiss SEM EVO 50 at 20 kV accelerating voltage. As-
received GNPs were dispersed in DMAc solvent in a ratio of around (1mg/20 mL)
using a water-bath ultra-sonicated instrument for 2 hours. After that, one drop of the
dispersion liquid was placed on silicon wafer substrates which had been carefully
stuck onto conductive carbon tape and in turn adhered to aluminium stubs.

The cryo-fractured surface of neat TPU-70 HS and TPU70/GNPs-NCs was studied


using a Zeiss SEM EVO 50 at an accelerated voltage of 20 kV as well. The
dumbbell-shaped samples were frozen in liquid nitrogen for 3-5 minutes then broken
using a gripper. The fractured samples were then stuck to the aluminium stubs with
carbon conductive tape (both of which were supplied by Agar Scientific) and coated
with approximately 30 nm of carbon using a coating machine (Edwards S150B
sputter coater).

For the TPU70/CF and TPU70/CF-MSCs, the samples were cut using a precision
cutting machine, cleaned thoroughly in water and left to dry at 50 °C overnight. The
cutting samples were placed onto aluminium stubs with carbon conductive tape and
coated with 30 nm of Au/Pd using a coating machine (Edwards S150B sputter
coater). In all samples except the GNPs, sliver paint was used to reduce the risk of
the samples becoming charged by forming a conductive bath between the samples
and the pin stubs, avoiding the accumulation of electrons on the samples.

The coating machine (Edwards S150B sputter coater) and both carbon fibre braid
that used to coating TPU-70 HS and TPU70-NCs and Au/Pd alloy film that used to
coat TPU70/CF and MSCs respectively are illustrated in Figure 3.21.

113
Coating Carbon fibre Au/Pd alloy
chamber braid film
Control
panel

(B) (D)

Sample
holder

Coated
Sample Coated
Sample
Vacuum
pump
(A) (C) (E)

Figure 3.21: A) coating machine set up, B) carbon fibre braid for coating neat TPU-
70 HS and TPU70-NCs and C) the holder of sample with the resultant coated
sample, D) Pu/Pd alloy film for coating TPU70/CF laminate and MSCs and E) the
sample holder with the resultant coated sample.

3.8.10 Transmission Electron Microscopy (TEM)

Background of TEM

TEM is a beneficial technique that can be used for studying the dispersion of
nanofiller in the polymer matrix and the effect of sonication on GNP flakes with high
resolution and high magnification. TEM uses an electron gun in the same way as an
SEM instrument, generating an electron beam that is transmitted to the sample after
passing through a condenser lens. It is then passed through an objective and project
lens, and finally, the image is obtained on a fluorescence screen [177, 178]. In
contrast to SEM, TEM is usually run at a highly accelerated voltage that reaches 200
kV. It is important to note that electron microscopy conducted at high voltage may
spoil the sample, owing to the microstructural defects that can be generated on the
specimen surface throughout observation [162, 176].

Sample Preparation for TEM

In this work, TEM was used to investigate the effect of sonication on GNP (M5, M15
and M25) flakes. Tests were carried out using a Philips CM200 TEM at 200 kV
accelerating voltage. The GNP samples were prepared in the same manner as the
SEM samples, except that the drops of suspension were dropped on Lacey carbon

114
films 300 Mesh on copper TEM grids (supplied by Agar Scientific) using a
micropipette. This grid was then placed in a fume cupboard overnight to allow the
solvent to evaporate.

3.8.11 Optical Microscopy

Optical microscopy was used to investigate the dispersion of GNPs in the TPU-70
HS matrix. For investigating the dispersion of GNPs, flattened films of TPU70-NCs
were fabricated using a hot press, cut into small pieces and placed on a glass slide.
The specimen was then examined using an Olympus BH-2 optical microscope fitted
with x10 and x50 objective lenses. AxioVision Rel. 4.8 software was used to capture
and adjust the optical microscopic images.

3.9 Rheological Characterisation

3.9.1 Background of Rheology

Rheology is defined as the science that describes the deformation and flow of matter
[179]. Rheology methods are widely used to study the effect of fillers on the
rheological behaviour of a polymer matrix. The viscoelastic properties of polymer
composites are necessary and important, as they are related to the processing,
microstructure and dynamics of composites [180]. The simplest model used to study
the rheological properties of thermoplastic polymers and polymer nanocomposites is
the parallel plate model. This model describes a Newtonian fluid, as illustrated in
Figure 3.22, where (h) represents the distance between two parallel plates filled with
inspected material. The bottom plate in this model is fixed, whereas the top plate,
which has a surface area A, is moved in the x-direction at a velocity V by an applied
force F. As a result, the fluid will be subjected to shear stress, , that is equal to F/A,
generated by applied force F. The magnitude of shear stress is proportional to the
𝑑𝑉𝑥
gradient of velocity that is developed between the plates, ; this can be written as
𝑑𝑦

follows:

𝑭 𝒅𝑽𝒙
= 𝝉 = [ ] ......... (3.39)
𝑨 𝒅𝒚

where  represents the viscosity of the fluid and is a parameter of its resistance to
flow.

115
V
Moving plate
h
Fixed plate

Figure 3.22: Schematic representation of Newtonian fluid for a parallel plate model
[1, 181]

The shear stress generated also causes the fluid to move at a shear strain (); this
𝑑𝑥
shear strain is equivalent to (𝑑𝑦). At any point in the gap h, the shear strain rate or

𝑑𝛾
shear rate can be defined as 𝛾̇ = , as shown below:
𝑑𝑡

𝒅𝒙 𝒅𝒙
𝒅𝑽𝒙 𝒅( ) 𝒅( ) 𝒅𝜸
𝒅𝒕 𝒅𝒚
= = = = 𝜸̇ ......... (3.40)
𝒅𝒚 𝒅𝒚 𝒅𝒕 𝒅𝒕

Therefore, equation (3.39) can be rewritten as follows:

𝝉 =  𝜸̇ ......... (3.41)
Equation (3.41) is known as Newton’s law of viscosity [182]. Most polymers exhibit
non-Newtonian behaviour, in which their viscosity is influenced by frequency () in
an oscillatory shear test and by shear rate in a steady shear test. In addition, the
nanofillers have a significant effect on the viscosity of a polymer when it is mixed
with them. In the case of non-Newtonian behaviour, the response of the dispersion
materials at a small strain is linear viscoelastic behaviour. This region is known as
the linear viscoelastic region (LVR), which refers to the network stability of melted
NCs before the sample is overstrained and prone to destruction.

To determine the LVR, the oscillation mode of the parallel plate rheometer was used.
In this mode, the sample was subjected to the small sinusoidal strain, , and will
respond with sinusoidal shear stress () at an angular frequency (). This shear stress
was shifted by a phase angle, ; therefore the loss modulus (G”) and shear storage
modulus (G’) are dependent on angular frequency (), as follows [183]:

𝜸 = 𝜸𝟎 𝒔𝒊𝒏 𝝎𝒕 ........ (3.42)

𝝉 = 𝝉𝟎 𝒔𝒊𝒏(𝝎𝒕 + 𝜹) = 𝑮′ (𝝎)𝒔𝒊𝒏 𝝎𝒕 + 𝑮′′ (𝝎)𝒄𝒐𝒔 𝝎𝒕 ......... (3.43)

116
where 𝛾0 and 𝜏0 represent the strain and stress before recovery state, respectively
[183].

Thus, the complex modulus (G *), shear rate 𝛾̇ and the complex viscosity (*) can be
written as [183]:
𝝉
𝑮∗ = 𝜸 = 𝑮 (𝝎) + 𝒊𝑮(𝝎) ........ (3.44)

𝒅𝜸
𝜸̇ = = 𝜸𝟎 𝒄𝒐𝒔 𝝎𝒕 ......... (3.45)
𝒅𝒕
𝝉
∗ = =  (𝝎) + 𝒊  (𝝎) ......... (3.46)
𝜸̇

3.9.2 Experimental Procedure of Rheological Test

In the present study, two types of rheological test were applied using a TA
Instruments HR-3 Discovery Rheometer to study the linear viscoelastic properties of
TPU-70 HS and TPU70/GNPs-NCs and to assess their processability. Both tests
were carried out using 25 mm-diameter aluminium parallel plates at 200 °C
separated by a distance of 1000 m. To determine the LVR of TPU-70 HS and its
nanocomposites in a melted state, an oscillatory strain sweep was accomplished
using a strain range from 0.01% to 1500% at a constant frequency of 1Hz. After the
LVR was identified, a frequency sweep was performed at a constant strain within the
LVR (0.2% in this work) and over a range of 0.1-300 rad/s. This test was carried out
to determine the shear storage modulus (G’), shear loss modulus (G”) and complex
viscosity (*) versus the angular frequency () for different GNP size loadings.

3.10 Impedance Spectroscopy

GNPs are known for their excellent electrical conductivity that reaches 107 S/m
parallel to a surface and 102 S/m perpendicular to a surface (see Table 3.6).
Impedance spectroscopy was used to measure electrical conductivity by applying an
alternating current (AC) to the samples and detecting their responses. In this study,
the impedance of NC matrices and multiscale composites (MSCs) were measured
using a system that consisted of a NumetriQ PSM1735 and an Impedance Analysis
Interface (IAI). The NumetriQ PSM1735 displayed the data and supplied the voltage,
whereas the IAI provided a wide range of shunts (1 m - 500 m) that allowed the
measurement of samples of various resistances. The Inductor-Capacitor-Resistor

117
(LCR) mode was used to obtain the parameters of frequency, inductance, impedance
(Z), capacitance, tan , phase angle () and Q factor. The electric resistance (R) can
be calculated using equation (3.47), which is shown below:

𝑹 = |𝒁| 𝒄𝒐𝒔 𝜽 ......... (3.47)

The test was performed using rectangular samples for both NCs (15 mm 10 mm  2
mm) and MSCs (15 mm  10 mm  4 mm). The dimensions of the samples were
very important because they were used to calculate the specific conductivity () of
the composites according to equation (3.48).
𝟏 𝒍
 = 𝑹 × 𝒃×𝒉 ......... (3.48)

where l is the length of the sample, and b and h are the width and thickness of the
sample respectively.

The samples were prepared using compression moulding for all samples of neat
TPU-70 HS and its nanocomposite, TPU70/CF laminate and MSCs. Before the test,
the samples were cut to rectangular shapes, and polished and coated using silver
paint in order to lessen the contact resistance between the contact surface of samples
and the conductive wires. The samples were then joined to the conductive copper
wires using silver-filled epoxy adhesive and hardener (supplied by Agar Scientific
Limited). The copper wires were connected to the samples at both ends and in
different directions in-plane and out-of-plane, as illustrated in Figure 3.23. To ensure
that the epoxy adhesive dried, the samples were placed in an oven overnight at 60
°C. The measurements were performed via a four-terminal Kelvin connection with a
Kelvin clip. A frequency range between 1 and 106 Hz was applied at voltage (rms)
amplitude of 1.0 V for all tests. Three samples were used for each type of composite
and each sample was tested twice. Before doing the test, the instrument was
calibrated using a known-value resistor to check that the instrument was working
accurately.

(A) (B)

Figure 3.23: Some electrical conductivity samples that were used during the present
work: A) in-plane direction and B) out-of-plane direction

118
3.11 Infrared (IR) Thermography

This technique was used to measure the thermal diffusivity of TPU70/CF laminate
and TPU70/CF-MSCs, owing to the difficulty of preparing samples of these
materials that would be fit for instruments used to measure thermal properties
without deteriorating the fabric alignment. This technique depends on measuring
thermal diffusivity () by heating one side of a sample using an external heat source.
The change in temperature can then be recorded using an IR camera set up in front of
the other side of the sample as shown in Figure 3.24.

Heat resources
Sample

IR Camera

Figure 3.24: Set-up of infrared thermography experiment


The thermal diffusivity was then calculated using equation (3.49) as described in the
flash method for thermal diffusivity in the ASTM E-1461 standard [184] and Parker
et al. [185]. The main benefit of this method lies in its ability to estimate dispersion
over a large area, reducing effort and cost, as well as measuring the thermal
properties of the system [186].
𝟐
𝜶 = 𝟎. 𝟏𝟑𝟑𝟖 𝑳 ⁄𝒕𝟏 ......... (3.49)
𝟐

where L is the thickness of the sample and 𝑡1 is the time required for the back surface
2

of the sample to reach half the maximum temperature. After that, a Matlab code was
used to allocate each pixel its own thermal diffusivity and could also be utilised to
map the quality of dispersion of the sample. The thermal conductivity could then be
calculated using equation (3.50) [184-186]:

119
𝑲 = 𝝆𝜶𝑪𝒑 ......... (3.50)

where K is the thermal conductivity,  is the density of the sample, and 𝐶𝑝 is the heat
capacity at constant pressure. The density of the sample was measured using a
Mettler Toledo machine. Before measuring the density, the sample was dried at 80°C
overnight; then the density was measured by measuring the weight of the samples in
air and in distilled water. Conversely, the heat capacity (𝐶𝑝 ) was measured via DSC
following the standard of Plastics — Differential scanning calorimetry (DSC) ISO
11357-4 [187] and [188]:

𝒎𝒔𝒂𝒑𝒑𝒉𝒊𝒓𝒆 ×𝜷  (𝑻)−𝒆𝒎𝒑𝒕𝒚 (𝑻)


𝑪𝒑 (𝑻) = 𝑪𝒑 𝒔𝒂𝒑𝒉𝒊𝒓𝒆 (𝑻) × × ( 𝒔𝒂𝒎𝒑𝒍𝒆 )......... (3.51)
𝒎𝒔𝒂𝒎𝒑𝒍𝒆 ×𝜷 𝒔𝒂𝒑𝒑𝒉𝒊𝒓𝒆 (𝑻)−𝒆𝒎𝒑𝒕𝒚 (𝑻)

Where: 𝑚𝑠𝑎𝑝𝑝ℎ𝑖𝑟𝑒 is the mass of the calibrated sapphire material, 𝑚𝑠𝑎𝑚𝑝𝑙𝑒 is the
mass of the sample in DSC,  is the scan rate (heating or cooling), and  represents
the heat flow rate for the sample, sapphire and blank material, respectively.

The samples used in this test were in square shape with 4 mm in thickness and 8 cm
for both width and length.

3.12 Analysis of Ultrasonic C-Scan Technique

Ultrasonic C-scan is a non-destructive testing (NDT) method that is widely used to


evaluate fibre-reinforced polymer composite materials and detect their various flaws
[189]. The C-scan technique can be conducted in two ways: the pulse-echo method
and the through-transmission approach. The former method depends on using one
probe which works as a transmitter and receiver at the same time, which is placed on
one side of the samples; a reflective plate is then placed on the other side.
Conversely, the through-transmission approach, which was adopted in the present
work, is considered the more conventional C-scan approach. This approach makes
use of two separate transducers: a transmitter and a receiver. They are placed on
either side of the sample and moved in a corresponding opposing path along it. Each
transmitter, sample and receiver is coupled with a water-jet to achieve the required
contacts. The signal amplitude is measured after travelling through different points in
the samples. The flaws within the sample will reflect the signals; as such, greater
reflection of the signal causes less signal to be transmitted through the sample [189,
190]. As a result, the strongest travelling signal will be detected and represented as a

120
map, showing the flaws and damaged areas in the sample. In general, a C-scan can
detect different types of defects, including porosity and delamination, disbanding,
and voids. Figure 3.25 illustrates the schematic of through-transmission ultrasonic C-
scan.

Water supply

Water-jet

Transmitter Receiver

Sample

Figure 3.25: Schematic of through-transmission ultrasonic C-scan [191]


In this work, a Midas NDT water-jet probe, through-transmission C-scan system was
used in conjunction with Zeus software v3.0. The composite samples were scanned
after being cut using a water-jet 0.5 MHz probe. A scan speed of 200 mm/s was used
to run the test and both the index step and grid sizes were 0.25 mm. After the impact
test, the samples were scanned again in order to detect the position of any damage.
Image-J software was used to calculate the damaged area due to the C-scan test
results, which were obtained as qualitative data rather than quantitative data. The
depth of the dent was measured using digital calibrators.

3.13 Acid Digestion Method

It is typically known that the formation of large voids within a composite during
manufacturing will reduce the properties of composite materials. On the other hand,
the fibres in composite materials are linked to many of their mechanical properties.
Therefore, it is crucial to calculate the volume fraction of fibre and the percentage of
voids in the composite. The acid digestion method was used in this work, in
accordance with ASTM D 3171-99, method I, procedure B [192]. This method is
used to measure CF volume fraction, as the oxidising polymer (burn-off matrix)

121
method at a temperature range of 500-600 C that is used to calculate glass-fibre
volume fraction is invalid for measuring CF/polymer matrices due to the likelihood
of CF oxidisation.

This method was carried out using at least three samples with a size of 1.5 ×1.5 cm2
or a weight not greater than 1 g, cut at random from each panel. The samples were
dried at 80 °C for 5 days before their density was measured. A Mettler Toledo
XP205 Delta Range electronic balance with a density measurement kit was utilised to
measure the weight of the samples in air and water, then their density was calculated.
After the density measurements had been recorded, the samples were taken and
immersed in 20 ml sulphuric acid within a glass tube and heated to 160 °C for two
hours. Drops of hydrogen peroxide were then added slowly to the dark solution until
the colour changed and became transparent. This change indicated that the TPU-70
HS matrix had been completely digested and oxidised. Vacuum filtration was then
used to filter the mixture and the residual CF reinforcement was washed with
distilled water several times until the pH = 7. Following that, the CF was placed in a
weighed crucible and dried in an oven at 120 °C overnight, then placed in a
desiccator to cool. Finally, the weight of the CF was recorded. In order to measure
the volume fraction for all of the constituents of the composite (fibre, matrix and
void), the following equation can be applied:
𝒎𝒇
⁄𝝆
𝑽𝒇 = ( 𝒎𝒄 𝒇 ) ........ (3.52)
⁄𝝆𝒄

where 𝑉𝑓 is the fibre volume fraction, and 𝑚𝑓 , 𝜌𝑓 , 𝑚𝑐 and 𝜌𝑐 are the mass and
density of the fibre and composite respectively. The volume fraction of the matrix
𝑉𝑚 can be calculated as follows:

(𝒎𝒄 −𝒎𝒇 )⁄
𝝆𝒎
𝑽𝒎 = ( 𝒎𝒄 ) ......... (3.53)
⁄𝝆𝒄

However, the composite consists of fibre, matrix and void; therefore, equation (3.54)
was used to calculate the void volume fraction(Vv) of the composite as a percentage.
𝑽𝒗 = 𝟏𝟎𝟎 − (𝑽𝒇 + 𝑽𝒎 ) ......... (3.54)

122
3.14 Mechanical Testing

3.14.1 Tensile Test

3.14.1.1 Tensile Test of Matrices


The tensile test of TPU-70 HS and TPU70/GNPs-NCs was carried out to determine
their tensile mechanical properties using a model 1122 Instron machine. This
machine was equipped with a 0.5 kN load cell and a distance between grips of 55
mm [155]. At least six samples were tested for each formulation, which was prepared
via injection-moulding according to ISO 527-2 type 1BA, as illustrated in Figure
3.26.

Figure 3.26: Tensile specimen (dumbbell-shaped specimen)[193]

The dimensions of the dumbbell-shaped specimens measured 30  0.5 mm in gauge


length, with a width of the narrow section of 5  0.5 mm, thickness of 2 mm or less
and an overall length 75 mm. Prior to testing, the samples were placed in a
conditioned room for 24 hours at a temperature of 25  1 °C and a relative humidity
of 502 %. The test was performed at a crosshead speed of 10 mm/min. Test results
were obtained as stress-strain curve and the modulus of elasticity (Young’s modulus,
E), yield strength, ultimate tensile strength and % elongation at break were recorded.
For example, the modulus of elasticity was determined from the initial linear region
or elastic region in a stress-strain curve that lay between 0-15% strains [194].

For TPU-70 HS and some of its nanocomposites, the stress-strain curve did not have
a clear point of yielding; a 0.2% offset method was therefore applied to determine the
yield strength value. The ultimate tensile strength was determined through the
maximum load point before failure or from the maximum point in the stress-strain
curve before rupture.

123
3.14.1.2 Tensile Test of TPU70/CF and Multiscale Composites
The tensile test specimens of TPU70/CF laminate composites and MSCs were
prepared according to ASTM D3039 in order to assess their tensile mechanical
properties [195]. Six specimens were tested for each laminate family (TPU70/CF,
TPU70/GNPsM5/CF, TPU70/GNPsM15/CF and TPU70/GNPsM25/CF). The
dimensions of the specimens were an overall length of 190 mm, width 15 mm and
thickness 2 mm with 50 mm in gauge length. The specimens were bonded at their
ends with end-tabs of glass fabric/epoxy composites (supplied by RS company) with
length 45 mm and 15 mm width using epoxy adhesive (supplied by RS company).

The strain measurement was carried out using a video extensometer with a high-
resolution digital camera that was used to track two opposing marks on the surface of
the specimens. Correction fluid was used to paint white marks on the dark surface of
the specimen, as illustrated in Figure 3.27.

Video Extensometer (VE)

Sample White marks-


dots

Figure 3.27: Set-up of a specimen in the tensile test machine with video extensometer
The strain of the specimen was measured by tracking the separation of the marks at
the start of the test (gauge length) and the current mark separation. In this technique,
tracking the marks could eliminate any error in strain measurement that might occur
due to mark stretching at high elongation. Before commencing each test, gauge
length was automatically measured and used in strain calculations [196]. The tests
were conducted on an Instron 5982 machine with a maximum load of 100 kN at a
constant crosshead speed of 2 mm/min. Bluehill 3 software was installed on the
machine computer and was integrated in the video extensometer camera. The
average tensile strength, elastic modulus and % elongation at break were then
obtained directly from this software on the tensile machine.

124
3.14.2 Flexural Test

This test was carried out on TPU70-NCs and MSCs in order to measure their flexural
properties using a three-point bending test according to ASTM D790-03 [197]. A
3344 Instron machine equipped with a 2 kN unit cell load was used to perform the
test at a crosshead speed of 2 mm/min. The width and thickness of the specimens of
both TPU70-NCs and MSCs were 12.7 mm and 2 mm respectively. Six specimens
were tested for each sample. The support span length (L) to depth (d) ratio was set at
16:1 for TPU70-NCs and 32:1 for MSCs, as recommended by the standard in
accordance with the structure of materials [197] as shown in Figure 3.28.

(A) (B)

Moveable Clamp

Sample

Figure 3.28: Flexural test set-up for A) TPU-70 HS nanocomposites and B)


multiscale composites
The diameter of the loading nose was 10 mm, while the diameter of the lower
supports was 6 mm. The calculation of flexural modulus (Ef) for both TPU70-NCs
and MSCs was accomplished using equation (3.55)

(𝑳𝟑 × 𝒁)
𝑬𝒇 = ⁄ ......... (3.55)
(𝟒 × 𝒃 × 𝒅𝟑 )
where L is the support span length; b is the width of the specimen; d is the
specimen’s thickness, and Z is the slope of the straight-line portion of the load-
deflection curve (N/mm). The magnitudes of the maximum flexural strength (f) of
TPU70-NCs and MSCs were estimated using equations and (3.56) and (3.57),
respectively.

𝟑𝑷𝑳
𝝈𝒇 = ......... (3.56)
𝟐𝒃𝒅𝟐

125
𝟑𝑷𝑳 𝑫 𝟐 𝒅𝑫
𝝈𝒇 = 𝟐𝒃𝒅𝟐 [𝟏 + 𝟔 ( 𝑳 ) − 𝟒 ( 𝑳𝟐 )] ......... (3.57)

where D is the deflection of the centre line of the specimen at the middle of the
support span (mm) and P is the load at a given point on the load-deflection curve
(N).

3.14.3 Interlaminar Shear Strength (ILSS)

A short-beam shear strength test (SBSS) is commonly used to determine the


interlaminar shear strength (ILSS) of composites owing to its simplicity, practicality
and the small sample size required [198, 199]. This test of TPU70/CF laminate and
MSCs was carried out according to the ASTM standard D 2344/D 2344M – 00 [200].
Three-point bending was used in an Instron 4301 machine equipped with a load cell
of 5 kN at a crosshead speed of 1 mm/min. As recommended by the standards [200],
the dimensions of specimens should be determined using their thickness, which was
4 mm in the present work. The span length (L) to thickness (h) ratio was considered
4:1 in accordance with the standard [200]. Therefore, the span length value was equal
to 16 mm. The width (b) and length of the specimen were calculated from the
relation (b/h=2) and (specimen length/h=6), respectively. Six specimens were tested
for each sample and the average value of ILSS was obtained. Before the test, the
specimens were conditioned for 24 hours at 23 °C and a humidity of 50%. Equation

(3.58) was used to calculate the value of ILSS (𝐹 𝑆𝐵𝑆 ):

𝟑𝑷𝒎𝒂𝒙
𝑭𝑺𝑩𝑺 = ......... (3.58)
𝟒𝒃𝒉

where 𝐹 𝑆𝐵𝑆 is the ILSS in MPa and Pmax is the maximum load that was observed
during the test (N). After the test, three different types of failure could be observed in
accordance with the standard. Figure 3.29 illustrates these three types, which include
interlaminar shear failure (where some of specimens failed in the present work),
tension or compressive failure, and finally inelastic failure.

126
Figure 3.29: Interlaminar shear strength (ILSS) failure modes: 1) interlaminar
shear, 2) tension or compression failure, and 3) inelastic deformation [200]

3.14.4 Mode-I Interlaminar Fracture Toughness

This test was carried out to measure the interlaminar fracture toughness (ILFT) of the
TPU70/CF laminates and MSCs using mode-I, double cantilever beam (DCB)
specimens as illustrated in Figure 3.30.

Figure 3.30: Double cantilever beam (DCB) specimen geometry [135]


The test conditions were performed according to the ASTM D5528-01 [201]. The
dimensions of specimens were 160 mm length and 20 mm wide and tabs made from
glass fabric/epoxy composites (the same one used for end-tabs tensile test) were cut
to the same dimensions of the specimens and glued to both sides of each specimen’s
surface in order to increase stiffness using 3M Scotch-weld. The overall thicknesses
of specimens were around 7 mm. The same idea was used in previous work by Doris

127
et al. [202]. Six specimens were used for each type of composites and MSCs. An
aluminium film with thickness around 13 m was used to form a pre-crack which
was inserted in the mid plane of each sample. The metal piano hinges were cut to the
specimen wide (20 mm) and fixed to the ends of the specimens where the inserted
film was started using the same adhesive (3M Scotch-weld). The beginning of the
crack, which represents the distance from loading points to the end of the inserted
film was 30 mm, as reported in previous work [47, 203]. A white spray painting was
used to paint the specimens and marked them on the side faces from the end of the
inserted film every 1 mm up to 60 mm, as shown in Figure 3.31.

Figure 3.31: DCB specimen of mode I interlaminar fracture toughness test specimen
An Instron 5969 with a 10 kN cell load was used to conduct the test at a crosshead
speed of 2 mm/min. A video extensometer camera was used to observe and follow
the propagation of the crack tip. A pip counter was used to record the load and
displacement by pressing it every 1 mm. In detail, firstly the specimen was loaded
and the crack propagation was recorded until 3 mm and the specimen then unloaded.
Secondly, the specimen was reloaded again without stopping and the pip counter was
pressed again after 3mm for each mm until the crack was reached 60 mm. The
modified beam theory (MBT) was applied to calculate the value of the critical energy
release rate of Mode I: 𝐺𝐼𝐶 [135, 201] using the equation below.

𝟑𝑷𝜹
𝑮𝑰𝑪 = ......... (3.59)
𝟐𝒃(𝒂+|∆|)

where P is applied load (N),  is load point displacement (mm), b is a specimen


width (mm), a is a crack length (mm) and |∆| crack tip rotation correction factor. The

128
value of |∆| is used because practically without it the value of GIC will be
overestimated since the beam is not perfectly built-in (namely, rotation may occur at
the front of delamination). The  can be calculated by generating a least squares plot
of the cubic root of compliance (C1/3) against the delamination length as shown in
Figure 3.32. The compliance (C) represents the ratio of applied load point
displacement to the applied load (/P).

Figure 3.32: schematic representative the calculation of  in MBT

3.14.5 Impact Resistance Test

An Instron CEAST 9350 drop weight impact machine equipped with CeastVIEW
5.94 3C software was used in this study to test the impact resistance of TPU70/CF
laminate and TPU70/CF/-MSCs. The test adopted an impact method introduced at
Queen Mary University of London (QMUL) by Prichard and Hogg [204]. This
method reduced the size of a specimen to 89 mm  55 mm instead of the dimensions
of 150 mm  100 mm that were recommended in the ASTM D7136 standard [205].
The reduction in sample size makes the test straightforward to conduct inside a
university laboratory and also reduces the amount of materials required to prepare
the test specimens. Figure 3.33 shows the set-up of the impact test introduced by the
QMUL method. In this test, a specimen was clamped between two plates with a
circular opening hole 40 mm in diameter. A hemispherical impactor with a 20 mm
diameter and a mass of 5.048 kg was used to cause damage to the specimen by
striking at three impact energy levels: 5 J, 15 J and 25 J. At each energy level, four
specimens were tested for each sample. In order to prevent a second strike that might
hit the sample after rebounding, the impactor was automatically held.

The results that were exported from the software were used to create graphs of force
versus time, energy versus time and force versus deformations. From those graphs, a
maximum contact force, maximum deflection and absorbed energy were obtained. A
digital depth gauge was used to measure the dent depth. Measurement of the dent
129
depth was carried out by calculating the difference between the centre of the dent and
four points at a distance of around 20 mm from each side of the depth centre.
Hemisphere 20 mm diameter

Specimen Diameter 40 mm

Figure 3.33: Schematic representation of impact test set-up [204]

3.14.6 Compression after Impact (CAI)

CAI is the most commonly used means of measuring the residual compressive
strength and assessing the damage tolerance of a laminated composite specimen after
being subjected to an impact test. A delamination is a form of internal damage that
would typically be generated in the specimen due to the impact test. It can either be
local or extend into a large area. As such, it will cause rapid and catastrophic
buckling failure when the specimen is subjected to compressive loading [206]. In this
study, the CAI test was carried out using an Instron 5989 equipped with a 600 kN
load cell and at a constant crosshead speed of 0.5 mm/min. An anti-buckling fixture
was used to fix the specimen and to prevent out-of-plane buckling failure as
illustrated in Figure 3.34.
P
Top loading
assembly

Test specimen

Slide support Slide support

Base assembly

Figure 3.34: Schematic representation of compression after impact (CAI) fixture


[207]

130
This figure represent Figure 3.34 illustrates the anti-buckling fixture and the sample
fixed with it [207]. For each sample, three specimens were tested which were the
same specimens used for the impact test in order to determine their residual
compressive strength. The average maximum compressive strength before and after
the impact test was obtained by the machine software (Bluehill 3) and the relation of
compressive force versus displacement was drawn as well.

131
Chapter 4: Effect of Graphene Nanoplatelet Size
and Annealing Treatment on the Properties of a
Matrix of TPU-70 Hard Segments

4.1 Introduction

This chapter focuses on studying the effect of incorporating three different types of
graphene nanoplatelets (GNPs) from XG SciencesTM (GNPM5, GNPM15 and
GNPM25 with 5, 15 and 25 m average particle diameter respectively) within TPU-
70 wt. %HS (TPU-70 HS) matrix on its dynamic, thermal, mechanical and electrical
properties. All of the nanocomposite (NC) materials in the present work were
prepared using an in-situ polymerisation technique. Early studies have found that in-
situ polymerisation is a more active method than other techniques used to prepare
nanocomposite (solution mixing and melt compound), which combines good
dispersion of nanofillers and strong interfacial interaction [48, 152-154]. As a result
of this finding and those of previous studies, which demonstrated that in-situ
polymerisation yielded better mechanical and electrical properties of TPU-70 HS
nanocomposites (TPU70-NCs) than other techniques [46], this technique was
adopted in this investigation.

The first part of this chapter has focused on the characterisation of GNP types using
SEM, TEM, and WAXD to scrutinise the effect of sonication on GNP flakes. The
second part studies the element composition on the surface of GNPs after extraction
from TPU-70 HS using CHNS and XPS techniques. The third part focuses on
studying the effect of annealing treatments and/or GNP size on the rheological
behaviour, thermal properties, conductivity behaviour, and mechanical properties of
nanocomposite matrices (NCMs). In addition, SEM was used to investigate the
fracture surface for neat TPU-70 HS and TPU70-NCs in order to scrutinise the
dispersion of xGNPs in the TPU-70 HS matrix. All of these characterisations were
carried out in order to give an understanding of the effect of the size of GNPs and
annealing treatment on the NC matrices’ performance and on the performance of
MSCs.

132
4.2 Characterisation of Graphene Nanoplatelets

4.2.1 Thermal Stability of GNPs

110

100
Weight Loss (%)
90

80

70 Graphite
GNPM5
60 GNPM15
GNPM25

50
0 200 400 600 800 1000
Temperature C

Figure 4.1: TGA curve of as-received graphite, GNPM5, GNPM15 and GNPM25
with average particle diameter 5m, 15 m and 25m respectively under a nitrogen
atmosphere
The thermal stability of as-received GNPs of different sizes was compared to that of
pure graphite under an N2 atmosphere, as shown in Figure 4.1. From this figure, the
graphite exhibits a steady thermal behaviour over a temperature range below 700 °C,
where it starts to decrease slightly. Meanwhile, the behaviour profiles of GNPM5,
M15 and M25 are alike, with minimal difference between them, and they are less
stable compared with graphite. At 800 °C, the percentage of mass loss for graphite,
GNPM5, GNPM15 and GNPM25 are approximately 3%, 10%, 8%, and 7% of their
initial weight, respectively. Similar results have been reported by Al-Shammari [208]
and Gedler et al. [209].

4.2.2 Characterisation of GNP Morphology

To study the effect of sonication on GNP flakes according to their shape or


exfoliation, the as-received GNPs were compared with sonicated ones using different
techniques, such as TEM, SEM and WAXD.

4.2.2.1 TEM and SEM Analysis


The conditions used in the sonication technique, such as time, frequency or
temperature, have a great effect on GNP size and exfoliation, as previously reported

133
by Nawaz et al. [210] and Khan et al. [211]. The researchers studied the effect of
sonication time on the length of the graphene flake, and revealed that sonication in a
solvent for an extended period reduced the length of GNPs from approximately 3 to
1m [210, 211]. Therefore, the present work adopted short times of around 2 hours
for sonication in order to prevent any damage to the flake shape or reduction in their
length. SEM, TEM and WAXD techniques were used to investigate the effect of
sonication on the GNPs (grade M, size 5, 15 and 25 m). Figure 4.2 shows SEM
images of GNPs before and after sonication.

A B

C D

E F

Figure 4.2: SEM images of GNPs before sonication A) GNPM5 (5m average
particle diameter), C) GNPM15 (15m average particle diameter) and E) GNPM25
(25m average particle diameter) and after sonication B) GNPM5, D) GNPM15 and
F) GNPM25. The arrows indicate the folded and rolled up GNPs
The figure clearly shows that no dramatic change has occurred in the shape of the
GNP flakes. This result may be related to the low mechanical forces generated

134
during sonication, which are not strong enough to exfoliate or damage the GNP
flakes because of the short time and lower frequency at which they were applied.
Figure 4.3 displays the TEM images of as-received and sonicated GNPs.

Figure 4.3: TEM images of GNPs before sonication: A) GNPM5, C) GNPM15 and
E) GNPM25; and after sonication: B) GNPM5, D) GNPM15 and F) GNPM25. The
arrows in the images indicate the clear wrinkling and folding areas of the different
GNP types
This figure indicates that the GNP flakes fold and roll up after sonication, which
means that there is no change in the GNPs’ shape; this finding is in agreement with
the SEM results. These presented results for both SEM and TEM images matched the
findings obtained by Al Bozahid [46]. The researcher’s study was describe in detail
in literature review (see sections 2.1.1.1 and 2.2.3) using the same TPU-70 HS as a
135
matrix and reinforced by GNPM15. It was found that the short time and frequency of
sonication process which are similar to the present work (frequency of 37 kHz and a
temperature of 30 C for approximately two hours) did not cause any further damage
to GNPM15 flakes when the flake still maintained their overlapping, folding and
rolling up, indicating the difficulty to attain a mono-layer GNPM15 [46].

4.2.2.2 WAXD Analysis


WAXD was carried out in order to examine the effect of sonication on the crystalline
structure of GNPs. Figure 4.4 shows the WAXD patterns of as-received GNPs (A-
GNPs) compared with sonicated GNPs (S-GNPs).

S-GNPM5 S-GNPM15 S-GNPM25


A-GNPM5 A-GNPM15 A-GNPM25

S-GNPM25
Intensity (a.u.)

A-GNPM25

S-GNPM15
A-GNPM15

S-GNPM5

A-GNPM5

10 20 30 40 50 60 70 80
2 (degree)
Figure 4.4: WAXD for as-received GNPs (M5, M15 and M25), donated by A-GNPs
(M5, M15 and M25) before sonication and compared to WAXD for S-GNPs (M5,
M15 and M25) after sonication
In this figure, no difference can be observed in the crystalline plane of as-received
and sonicated GNPs. All of the diffractogram patterns of GNPs display a very
intense and narrow peak at 26.6, corresponding to an interlayer spacing of 3.35 Å
and the X-ray reflection plane (002). This peak is approximately the same to the
reported value of 26.6  and 3.35 Å interlayer spacing of GNP, as reported by Geng
et al. [28] and by Chandrasekaran et al. [27] . In addition, there is a small peak
corresponding to the X-ray reflection plane (004) at 55, representing an interlayer
spacing of 1.67 Å. Both 55 and an interlayer spacing of 1.67 Å are close to the
reported value of 54.7 and an interlayer spacing of 1.68 Å by Caradonna et al. [212]
depending on different types of GNPs. In summary, after sonication, the location of

136
the highest and lowest peaks remains the same. Furthermore, there is no peak
observed at low value of 2, which reveals that no change occurs in the interlayer
distance as a result of the exfoliation of individual layers of GNPs. The same result
was found by Spitalsky et al. [213] who confirmed no observation of peaks at low
value of 2 of sonicating expanded graphite nanoplatelets in a toluene solution for 20
minutes. This sonicated mixture was used to reinforce Styrene-Isoprene-Styrene
copolymer by solution mixing process.

The current results of WAXD of as-received and sonicated GNPs are compatible
with the previous TEM and SEM characterisation of GNPs before and after
sonication as well (see section 4.2.2.1).

4.3 TPU-70 HS NCs Characterisation

With the addition of GNPs, the structure and formation of TPU-70 HS might change
depending on the size and concentration level of nanoparticles. In this section, the
effect of GNP size and annealing treatment on NC matrices is studied, in order to
provide information on their nature so as to better work with them during the
preparation of multiscale composites (MSCs).

4.3.1 CHNS and XPS analysis

CHNS and XPS analyses were carried out in order to characterise the interfacial
interaction between GNPs and TPU-70 HS. This objective can be met by extracting
GNPs from TPU-70 HS and comparing the composition of their elements, such as C,
N or another element, with that of the as-received samples. The GNPs were extracted
from the TPU70/NC solution using a centrifugal machine for 45 minutes at a speed
of 1500 rpm. The nanoparticles were then washed a second time under the same
conditions. Both the GNPs extracted from the first and second wash were dried at
80C for 5 days; the resulting GNP powder was used for CHNS and XPS analysis in
order to compare them. Table 4.1 shows the CHNS elemental analysis results for as-
received GNPs and those extracted from TPU-70 HS nanocomposites using DMAc
solvent. This test shows that the as-received GNPs were contaminated with sulphur,
the amount of which decreases after the first and second wash. After the first wash,
the amounts of hydrogen and nitrogen on the GNPs’ surface increase, while the
amount of C decreases. Meanwhile, in the second wash, nitrogen and hydrogen

137
amounts decrease, whereas C increases and this behaviour will be explained latter in
this section.
Table 4.1: Atomic concentration of GNPs (M5, M15 and M25) as measured by
CHNS elemental analysis
Element (%)
Sample
C H N S
GNPM5- As received 90.20 0.35 0.18 0.56
First Washing 73.64 5.33 4.36 0.30
Second Washing 90.23 0.99 1.14 0.54
GNPM15-As received 93.09 0.26 - 0.5
First Washing 83.46 2.94 2.55 0.20
Second Washing 90.36 1.18 1.16 0.35
GNPM25- As received 91.79 0.28 0.71 0.61
First Washing 82.90 3.03 2.68 0.31
Second Washing 86.86 1.69 1.67 0.42

Figure 4.5 exhibits the low-resolution XPS survey scan of three types of GNP
nanoparticles (M5, M15 and M25) at three different situations: as-received, after first
wash, and after second wash, respectively.

(A) (B)
GNPs-As received GNPs-First washing
C1s O1s
C1s
Intensity (a.u.)

N1s O1s N1s


S2P
GNPM25 GNPM25

GNPM15 GNPM15

GNPM5 GNPM5

0 100 200 300 400 500 600 700 800 900 0 100 200 300 400 500 600 700 800 900
Binding Energy (eV) Binding Energy (eV)
(C)
GNPs-Second washing
C1s

O1s
Intensity (a.u.)

N1s

GNPM25

GNPM15

GNPM5

0 100 200 300 400 500 600 700 800 900


Binding Energy (eV)

Figure 4.5: Low-resolution XPS survey scan of GNP types (M5, M15, and M25): A)
as received; B) after first wash; C) after second wash

138
Figure 4.5 A shows that all samples of as-received GNPs have the same C1s and O1s
peaks, which are located at  281.3 eV and  531 eV, respectively; these are roughly
the same as the reported results by Junid [10] and Dul et al. [214]. For example,
Junid [10] reported that the main peaks of C1s and O1s of as received GNPM25 were
located at 284 eV and 533 eV. Dul et al. [214] found that the main peaks of C1s
and O1s of as received GNPM5 were located at 285 eV and 231 eV. This small
difference between the present study and Junid [10] and Dul et al. [214] studies may
attribute to the difference in production of GNPs which are unfortunately unknown.
There is also a small amount of nitrogen (approximately 0.3% to 0.5%) detected,
with a binding energy of 400 eV [214]. These results demonstrate that all of the GNP
types belong to the same surface functionality group, which produces similar
interactions with the TPU-70 HS matrix

Figure 4.5 B after, the first wash, the intensity of O1s and N1s peaks rises and the
total oxygen content reaches 16 % for all samples of GNPs. This outcome can be
attributed to the increase in the interaction between carbon and oxygen, to form more
functionality groups, such as carbonyl groups (C=O) and carboxylate groups (O-
C=O) [214-216]. Meanwhile, the content of nitrogen increases to reach a range of
approximately 3-4 % due to the presence of cyanate groups (NCO) resulting from the
TPU structure [48]. Therefore, the atomic percentage concentration of carbon
decreases after the first wash, as summarised in Table 4.2.

Table 4.2: Atomic concentration of GNPs (M5, M15 and M25) as measured by XPS
analysis

Element (%)
Sample
C1s O1s N1s S2p
GNPM5 - As received 94.87 4.42 0.32 0.38
First Wash 80.08 15.88 4.04 -
Second Wash 82.38 13.80 3.82 -
GNPM15 -As received 94.66 4.87 0.42 0.04
First Wash 80.57 15.67 3.77 -
Second Wash 83.75 12.84 3.41 -
GNPM25 - As received 94.76 4.67 0.48 0.09
First Wash 79.93 16.05 4.02 -
Second Wash 83.34 13.51 3.51 -
As shown in Figure 4.5 C, the highest of the peaks of O1s and N1s after the second
wash slightly decrease, indicating a reduction in the content of O and N respectively,

139
with a slight increase in C content. These results suggest that some of the groups
containing oxygen or nitrogen were removed after the second wash.

To conclude, both the CHNS and XPS results are compatible, in spite of a minor
reduction in the content of N and C. These results indicate the interaction between
the GNPs’ surface functionality and isocyanate, urethane, hydroxyl or ether groups in
the TPU70-HS structure, which causes an increase in nitrogen or oxygen and a
reduction in C after the first wash. However, in the second wash, some of these
interactions were broken and removed, causing an increase in the amount of C and a
decrease in the amounts of nitrogen and oxygen. These results indicate that some of
these interactions are physical, such as hydrogen or Van der Waals bonds, and can
therefore easily be removed by washing with solvent. Conversely, the other
interfacial interactions are chemically crosslinked; this outcome indicates that they
are likely to be covalent bonds that withstood the second wash and were not easily
removed.

Finally, both the CHNS elemental analysis and XPS technique detected a small
amount of sulphur in the as-received GNPs, as illustrated in Table 4.1and Table 4.2,
respectively. These amounts can be attributed to residual sulphur on the surface of
the as-received GNPs as a result of the manufacturing process. On the other hand, the
type of manufacturing process used is still unclear, and the sulphur amounts may be
due to acid treatment. Unfortunately, it is not possible to know what type of acid was
used in this treatment (taking into consideration that the datasheet for the GNPs
supplied by XG Sciences mentioned that the GNPs contained less than 0.5% sulphur)
[150]. Overall, after the first and second washes, the XPS technique could not detect
any sulphur in any of the GNP samples, owing to their level being lower than the
XPS level.

4.3.2 Gel Permeation Chromatography (GPC)

In the present work, the GPC technique was used to measure the number-average
molecular ̅𝑛 ,
weight 𝑀 the weight-average molecular ̅𝑤 ,
weight 𝑀 and the
polydispersity index (PDI) of neat TPU-70 HS samples. In order to study the effect
of GNPs on the PDI of a TPU-70 HS matrix, the GNPs were extracted from a TPU-
70 HS matrix owing to the practical difficulties of measuring Mw for the NCs.
Besides, the GNPs were extracted to prevent them from potentially entering the

140
chromatography separation column. After extracting GNPs using a centrifuge and
solvent DMAc, the TPU-70 HS was dried at 80°C for three days. The following day,
the TPU-70 HS was dissolved in THF in the same manner mentioned in chapter three
̅𝑛 and 𝑀
(see section 3.8.1). Figure 4.6 illustrates the effect of GNPs sizes on the 𝑀 ̅𝑤
of neat TPU-70 HS and its NCs and Table 4.3 summarises the̅̅̅
𝑀𝑛 ,̅̅̅
𝑀𝑤 , and PDI
results for neat TPU-70 HS and extracted ones from NCs.

Mn (g/mol) 3.5x104
1.8x104
Mw (g/mol)
4
3.0x104
1.5x10

2.5x104

Mw (g/mol)
Mn (g/mol)

1.2x104
2.0x104
3
9.0x10
1.5x104

6.0x103
1.0x104

3.0x103 5.0x103

NPM
5 M15 M25
0 HS . %G % GNP % GNP
-7 wt wt. wt.
TPU 70/5 70/5 70/5
TPU TPU TPU

Figure 4.6: 𝑀̅𝑛 and 𝑀


̅𝑤 for neat TPU-70 HS and extracted TPU-70 HS matrix from
its nanocomposites obtained from Gel permeation chromatograph test.
Table 4.3: Gel permeation chromatograph for neat TPU-70 HS and extracted TPU-
70 HS matrix from its nanocomposites

Sample ̅𝑛 (g/mol)
𝑀 ̅𝑤 (g/mol)
𝑀 PDI
Neat TPU-70 HS 11868  403 27893  297 2.40  0.08
TPU-70/5wt%GNPM5-NCs 13350  257 25550  195 2.10  0.02
TPU-70/5wt%GNPM15-NCs 13445  170 26747  285 2.03  0.01
TPU-70/5wt%GNPM25-NCs 15580  172 31424  68 2.01  0.02
The presences of GNP flakes in the TPU-70 HS matrix after extraction do not cause
a dramatic change in PDI, which indicates the equimolar ratio of the polyol and
isocyanate. This result also indicates that the GNPs did not cause any chain scission
during the preparation of TPU-70 HS nanocomposites [217, 218]. The increase in the
̅𝑤 of TPU70/GNPM25 (as illustrated in Figure 4.6) in comparison with
value of 𝑀
that of neat TPU-70 HS and other NCs indicate the effect of high surface area of
GNPs to create physical interactions with cyanate groups of HS in TPU structure and

141
then help to increase the length of molecular chain of TPU. These results confirm the
previous results of CHNS and XPS (see section 4.3.1) that demonstrated an increase
in the amount of concentration of O and N of as-received GNPs after incorporating
with TPU-70 HS due to the formation of physical cross links between functional
groups of GNPs and cyanate and/or carbonyl groups of TPU-70 HS.

4.3.3 Structure and Morphology Characterisation

The influence of the dispersion of GNPs and their size on the morphology of the
resultant TPU70/GNPs-NCs was investigated using SEM, optical microscopic (OM),
and WAXD techniques to discover the relationship between the structure and
properties of this system.

4.3.3.1 SEM Analysis


The dispersion state of nanofillers and the morphology of the TPU-70 HS have a
significant influence on the performance of NCs. Poor distribution or dispersion of
nanofillers within a polymer matrix can affect the polymer morphology and the
properties in a highly detrimental way, producing composites with undesirable
properties. The SEM technique was performed in order to give a straightforward
perspective on the dispersion of various GNP sizes within TPU-70 HS. Figure 4.7
shows the SEM images of a cryogenic fracture surface for TPU-70 HS and
TPU70/GNPs-NCs. Images on the right-hand side of this figure represent highly
magnified versions of the left-hand images, which are in low magnification; the
yellow arrows in the magnified images indicate the GNP flakes. In this figure, the
neat TPU-70 HS displays a smooth fracture surface which becomes rough with the
addition of 5wt% GNPs. This outcome may indicate the possibility of a crack
passing through the GNP flakes, causing a brittle fracture on the surface of the
TPU70-NCs. Furthermore, all the NC images show good dispersion of the GNPs
within the TPU-70 HS matrix. This observation contradicts with the results of a
previous study [219], which revealed that GNPM25 gives better distribution and a
rougher surface than GNPM5 due to its high surface area. This finding could be
assigned to the difference in methods that were used to produce NCs and the type of
polymer matrix, as the latter used melt compounds and polyamide 6 to produce
polyamide 6/GNPs-NCs.

142
Figure 4.7: SEM images of cryogenic-fracture surfaces of A) neat TPU-70 HS, B)
TPU70-NCs incorporated with GNPM5, C) GNPM15, and D) GNPM25 at low and
high magnifications. The yellow rectangular area in the low magnified images on the
left-hand side indicates the magnified region on the right-hand side. The yellow
arrows in the highly magnified images indicate the GNPs flakes, while the red curves
indicate the path of electrons due to form conductive network of GNPM15 and
GNPM25

143
Previous studies have confirmed that at low loading, carbon nanofillers are separated
from each other in the polymer matrix and become very close at high loading,
leading to agglomeration [220-222]. When agglomeration takes place at high
loading, in this situation it establishes a conductive network that helps to transport
electrons and improve electrical conductivity. However, the similarity in the rough
fracture surface of GNPM5, M15 and M25 NCs confirmed the suggestion that a
network for electrical conductivity was formed at 5wt% for GNPM15 and M25 due
to their large size as illustrated in Figure 4.7 C and D respectively. Such similarity
contrasts with other suggestions that GNPM5 requires more than this amount. In
addition, a previous study on GNPM25-NCs [219] observed folding and rolling of
GNPs flakes within a polymer matrix structure, while the current images do not
indicate that, or perhaps it is merely difficult to discern in the present images.

4.3.3.2 Optical Microscopic (OM) Analysis


Optical microscopic (OM) analysis was another technique carried out in order to
investigate the dispersion of GNPs within the TPU-70 HS matrix. The OM technique
is a very simple method used to examine GNP distribution, which is represented by
bright regions. Figure 4.8 illustrates the optical images of the TPU70/GNPs-NCs
based on various GNP sizes.

A B

200 m
200 m

200 m

Figure 4.8: Optical micrograph images of the TPU70/GNPs-NCs based on different


GNP sizes: A) GNPM5, B) GNPM15, and C) GNPM25, respectively

144
Unlike in the SEM images, it is very easy to recognise the agglomeration regions,
which are represented by glowing areas in OM images in Figure 4.8. As such, in this
figure, GNPM15 and GNPM25 exhibit more glowing images compared to GNPM5
due to their high surface area. More glowing regions can be attributed to some
agglomerations of GNPs that form due to their ability to restack. The agglomeration
itself can affect the uniform distribution of the nanofiller. In addition, this
agglomeration increases with increased loading of nanofillers, as reported by
previous studies [46, 223, 224].

4.3.3.3 WAXD Analysis


The mechanical properties of semicrystalline thermoplastic are strongly influenced
by their microstructure, which is described by the degree of crystallinity, crystal size
and orientations. Their crystal structure and crystallinity can be assessed using
WAXD experiments, where the former depends on thermal history, the presence of
possible nucleation positions and manufacturing process [114].

The WAXD patterns for TPU-70 HS and TPU70/GNPs-NCs before and after
annealing are shown in Figure 4.9 A and B, respectively. From this figure, the
WAXD patterns for neat TPU-70 HS before annealing show a strong peak located at
around 2 = 19.16 and d-spacing around 4.65 Å. This peak relates to the existence
of a short-range ordering structure in the hard domain (HD). The other peak
represents a halo peak and belongs to the amorphous phase (soft phase (SP)) [46,
225].

The XRD patterns of TPU70/GNPs-NCs display both peaks provided by GNPs,


whilst the peak of neat TPU-70 HS disappears. This indicates that the addition of
GNPs, regardless to their size, affects the semi-crystalline structure of TPU-70 HS.
Otherwise, the peak signal of TPU-70 HS probably does not really disappear but is
just a smaller amount of material and thus a weaker signal would be detected in
comparison with peaks of GNPs. As demonstrated previously, the increase in the
loading of GNPs dispersion allowed the graphene sheet to stack together, leading to
the increase in the peak intensity of polymer-NCs [24, 226]. Since this study uses
approximately a large amount of GNPs, the disappearance of HS crystalline peak
might be attributed to the preceding reason.

145
A) Non-Annealed B) Annealed
TPU-70HS
TPU-70HS
TPU70/5wt%GNPM5 TPU70/5wt%GNPM5
TPU70/5wt%GNPM15 TPU70/5wt%GNPM15
TPU70/5wt%GNPM25
TPU70/5wt%GNPM25
Intensity (a.u.)

5 10 15 20 25 30 35 40 45 50 55 60 5 10 15 20 25 30 35 40 45 50 55 60
2 () 2 ()

Figure 4.9: WAXD patterns of neat TPU-70 HS and TPU70-NCs incorporated with
5wt.% of GNPM5, GNPM15 and GNPM25 A) before annealing and B) after
annealing
From Figure 4.9 A and B, the small scattering peak of TPU70/GNPs-NCs
corresponds to 2 = 55, which is the same place where the scattering pattern of pure
crystalline GNPs is observed (see the peak in Figure 4.4). Similarly, the sharp peak
of all TPU70/GNPs-NCs is in the same place as that of the pure GNPs located at 2
= 26.6, meaning that no shift in the peak of GNPs is observed. The GNPs remain in
their crystalline plane, indicating that there is no dramatic change in the GNPs’
crystallography structure which can cause any variation in its magnitude [46].
Bandla and Hanan [227] reported the same results for the WAXD of GNP NCs when
they observed no shift in the peak of GNPs of reinforced polyethylene terephthalate
(PET). Their conclusions confirmed that the GNPs were not noticeably exfoliated,
and results in strong support for the results presented here in spite of the differences
in the structure and behaviour of their matrices. Although Bandla and Hanan’s [227]
study and the present one are used different thermoplastic matrices, they used similar
nanofillers reinforcements (GNPs). Therefore, Bandla and Hanan [227] results can
be used to compare with present result in order to give an indication to the effect of
incorporating GNPs in thermoplastic matrices on the behaviour of resultant
nanocomposites.

Furthermore, the sharp peak of the TPU70/GNPs-NCs became broader, and its
intensity increases with the addition of GNPM15 and M25 compared to GNPM5, due

146
to their higher surface area [228-230]. The same results were described by Halit
[219].

However, for TPU70/GNPs-NCs (Figure 4.9 B), the annealing treatment did not
show any significant change in NCs when their sharp peak and intensity showed
slight differences for GNPM15 and M25, while GNPM5 exhibited a slight increase
in its intensity. This emphasised the ability of GNPs to prevent HS within the TPU
matrix from reordering, which then affects its crystalline phase. On the contrary, the
annealing treatment slightly increases the sharp peak and its intensity for neat TPU,
which can be attributed to the increase in the crystalline phase (HD) compared with
the amorphous phase (SP).

4.3.4 Thermal Analysis

4.3.4.1 TGA
Thermogravimetric analysis (TGA) was carried out to observe the effect of GNPs on
the thermal stability of TPU-70 HS and its nanocomposites under a nitrogen
atmosphere. Figure 4.10 A and B illustrates the TGA and DTG thermogram results
for TPU-70 HS and its NCs. The onset temperature (Tdonset) of degradations for neat
TPU and its NCs was considered as the temperature corresponding to 5% mass loss
[231].

The DTG curves of TPU-70 HS and its NCs exhibit two to three stages of the
thermal decomposition pattern. The first one corresponds to the thermal
decomposition of the hard segments (HS) and is detected in the range of 280-340 °C,
while the second one represents the thermal decomposition of the soft segment (SS),
as reported previously [232-234], and takes place in the range of 340-410 °C [46, 54,
235, 236]. The first stage involves the degradation of urethane linkage converting to
the isocyanate and alcohol [237], whereas the second stage relates to the greatest
mass loss of soft segments. By adding GNP, these temperatures are shifted to a
higher value. The third stage, which is detected in neat TPU-70 HS at a range of 450-
600 °C, is connected to the urea decomposition and other stable isocyanate
derivatives [54, 235, 236]. The Tdonset, residue mass percentage at 600 °C and the
maximum rate of degradation temperatures (Tdmax1 and Tdmax2) are tabulated in Table
4.4, as acquired from DTG thermograms of TPU-70 HS and its NCs.

147
(A)
105 100

98
90
96
5% mass Loss

Weight Loss (%)


75
94
100 150 200 250 300
60
TPU-70HS
45 TPU70/5wt. %GNPM5
TPU70/5wt.%GNPM15
30 TPU70/5wt. %GNPM25

15

0
0 100 200 300 400 500 600 700 800
(B)
1.8
TPU-70HS
TPU70/5wt. %GNPM5
Derivative Weight (%/C)

1.5 TPU70/5wt.%GNPM15
TPU70/5wt. %GNPM25
1.2

0.9

0.6

0.3

0.0

0 100 200 300 400 500 600 700 800


Temperature (C)

Figure 4.10: A) TGA and B) DTG thermograms for TPU-70 HS and TPU70/GNP
nanocomposites based on various GNP sizes
Table 4.4: TGA and DTG results for TPU-70 HS and TPU70-NCs incorporated with
5wt.% of GNPM5, GNPM15 and GNPM25 with average particle diameter 5, 15 and
25m respectively
Residual Mass %
Samples Tdonset Tdmax1 Tdmax2
at 600 °C

TPU-70 HS 293 ± 3 325 ± 1.0 365  3.0 7.0 ± 0.2


TPU70/5wt%GNPM5-NCs 294 ± 2 339 ± 3.0 378  2.0 9.0 ± 0.5
TPU70/5wt%GNPM15-NCs 304 ± 2 345 ± 1.0 381  0.5 10.0 ± 0.5
TPU70/5wt%GNPM25-NCs 304 ± 3 346 ± 1.5 382  1.0 12.0 ± 0.5

148
In general, all TGA curves show a marginal improvement in the thermal stability of
TPU and its nanocomposites. However, the Tdonset of TPU-70 HS (shown in Figure
4.10 A and Table 4.4) is slightly shifted towards a higher temperature by the addition
of GNPs. It corresponds to 293 °C for TPU-70 HS, which increases with the addition
of GNPs to reach 294 °C, 304 °C, and 304 °C for TPU70/GNPM5, M15 and M25-
NCs, respectively. Furthermore, the residue mass at 600 °C for neat TPU-70 HS was
recorded to be 7%, which rises slightly with the addition of GNPs to be 9 %, 10 %,
and 12 % for GNPM5, M15 and M25 NCs samples, respectively. This improvement
in the thermal stability of neat TPU-70 HS with the addition of GNPs can give
reasonable evidence that GNPs act as a barrier within the TPU matrix, leading to a
delay in the production of volatile products during decomposition or hindering their
diffusivity due to the formation of a tortuous path [234, 238-240]. The same results
were reported by Zhang et al., which give good agreement with this work [222].
However, the improvement in thermal stability produced by GNPM25 and GNPM15
was high compared to that observed for the GNPM5 samples. It can be deduced that
the GNPM25 and GNPM15 provide the best thermal stability for the TPU-70 HS
matrix in comparison with GNPM5 at the same loading due to their high surface area
and aspect ratio.

The highest peak in the DTG curves (shown in Figure 4.10 B) represents the
maximum rate of degradation temperature (Tdmax1) for HS. This temperature rises
with the presence of GNPs to reach approximately 339  3, 345  1, and 346  2 for
TPU70/GNPM5, M15 and M25/NC samples respectively, compared to 325  1 for
neat TPU-70 HS. Meanwhile, the second peak in DTG represents the maximum rate
of degradation temperature (Tdmax2) of the SS, which begins to degrade during the
second stage. The Tdmax2 of SS increases from 365  3 °C of neat TPU-70 HS to
become 378  2 °C, 381  0.5 °C, and 382  1 °C for TPU70/GNPM5, M15 and
M25-NCs, respectively. In general, Tdmax for HS and SS is improved by the addition
of GNPs in comparison with the neat TPU-70 HS samples. It is supposed that the
increase in the maximum rate of degradation temperatures in the first and second
stages may be related to the thermal stability of GNPs and its interfacial interaction
with the TPU-70 HS matrix, as reported previously [50, 241]. The good dispersion of
GNPs in the TPU matrix and the strong interfacial interaction between the GNPs and

149
TPU-70 HS enhance the intermolecular interaction of TPU-70 HS molecular chains
and delay their decomposition at high temperature [241-243].

4.3.4.2 DSC Analysis


DSC experiments were performed to investigate the effect of the addition of GNPs
and their size and annealing treatment on the crystallisation and melting temperature
behaviour of TPU-70 HS and TPU70/GNP/NCs. The thermograph curves of first
heating, cooling and second heating DSC cycles for TPU-70 HS and TPU
nanocomposite before and after annealing are plotted in Figure 4.11. In this figure,
A, C and E curves represent DSC scans before annealing and B, D and F curves
represent DSC scan after annealing. The quantitative results of melting temperature
(Tm), changes in heat of fusion (Hm), degree of crystallinity (XC), crystallisation
temperature of HS (TCHS) and crystallisation enthalpy (HC) of TPU-70 HS and its
various NCs are summarised in Table 4.5, Table 4.6, and Table 4.7 respectively.

It is reported in previous studies that the HS in the TPU structure, which represents
the crystalline phase, tends to reorder itself at a temperature above its glass transition
temperature during annealing treatment [72, 244]. For this reason, the annealing
treatment during this investigation was carried out at 80 °C for three days, which was
considered the best period for annealing in order to improve crystallinity and obtain
better properties of TPU-70 HS, as reported by Al-Obad [47].

In general, the melting temperature (Tm) of TPU-70 HS from the first heating before
and after annealing, as shown in Figure 4.11 A and B, respectively is higher than its
counterpart from the second heating, as shown in Figure 4.11 E and F, while the Tm
of the NCs is almost unchanged. This discrepancy was ascribed to the thermal
history of the neat polymer related to the manufacturing process, which causes this
difference [231].

The highest endothermic peak in the first and second heating cycle represents the
melting temperature (Tm) of the microcrystalline hard domain (HD) as illustrated in
Figure 4.11 A, B, E and F, which consists of a long ordering of HS of TPU-70 HS. It
can be observed that the Tm of neat TPU-70 HS shifts to a higher temperature in the
second heating cycle and before annealing treatment with the addition of GNPs,
regardless of the size of the GNPs (refer to Table 4.5 and Table 4.7). Meanwhile, the

150
value of heat of fusion associated with the Tm of neat TPU-70 HS is almost
unchanged upon the incorporation of GNPs.

(A) (B)
1st-heating Scan-Non-annealed 1st-heating Cycle-Annealed
TPU70/5wt.%GNPM25
TPU70/5wt.%GNPM25

Exo
Exo

TPU70/5wt.%GNPM15
TPU70/5wt.%GNPM15
TPU70/5wt.%GNPM5
TPU70/5wt.%GNPM5
TPU-70HS
TPU-70HS
Heat Flow (W/g)

Tm

Heat Flow (W/g)


Tm
TgHP TgHP

Tm

Tm Tm

25 50 75 100 125 150 175 200 25 50 75 100 125 150 175 200
(C) (D)
Cooling Scan-Non-annealed Cooling Scan-Annealed
TPU70/5wt.%GNPM25
TPU70/5wt.%GNPM25
Exo
Exo

TPU70/5wt.%GNPM15 Tc TPU70/5wt.%GNPM15
TPU70/5wt.%GNPM5 Tc
TPU70/5wt.%GNPM5
TPU-70HS
TPU-70HS
Heat Flow (W/g)
Heat Flow (W/g)

Tc

25 50 75 100 125 150 175 200 25 50 75 100 125 150 175 200
(E) (F)
2nd-heating Scan-Non-annealed 2nd-heating Scan Annealed
Exo

TPU70/5wt.%GNPM25 TPU70/5wt.%GNPM25
Exo

TPU70/5wt.%GNPM15 TPU70/5wt.%GNPM15
TPU70/5wt.%GNPM5 TPU70/5wt.%GNPM5
TPU-70HS TPU-70HS

Tm
Heat Flow (W/g)

Heat Flow (W/g)

Tm

Tm

25 50 75 100 125 150 175 200 25 50 75 100 125 150 175 200
Temperature(C) Temperature(C)

Figure 4.11: DSC curves for TPU-70 HS and TPU70-NCs incorporated with 5 wt.%
of GNPM5, GNPM15 and GNPM25 with average particle diameter 5, 15 and 25 m
respectively at 10° C/min heating/cooling rate in order to show the effect of GNP
sizes and annealing treatment on their melting temperature, glass transition
temperature and crystalline temperature. Where A), C) and E) are DSC scan cycles
before annealing and B), D) and F) are DSC scan cycles after annealing

151
Table 4.5: DSC data of first heating cycle for TPU-70 HS and TPU70/GNPs-NCs
before and after annealing
Before Annealing After Annealing
Sample
Tm1 Hm1 Xc1% Tm1 Hm1 Xc1%

TPU-70 HS 184.0± 0.4 24.8 ± 0.5 14.0 0.2 186.0 ± 0.1 22.7 ± 0.4 13.0 0.2

TPU70/GNPM5-NCs 182.0 ± 1.2 24.4 ± 0.6 15.1 0.3 184.0 ± 0.1 24.7 ± 0.1 15.0  0.2

TPU70/GNPM15-NCs 184.0 ± 0.7 24.0 ± 0.6 14.6 0.3 185.0 ± 0.4 23.8 ± 0.8 14.4  0.3

TPU70/GNPM25-NCs 185.0 ± 0.4 25.0 ± 0.3 15.2  0.1 186.0 ± 0.1 25 ± 0.8 15.0  0.4

Table 4.6: DSC data of cooling cycle for TPU-70 HS and TPU70/GNPs-NCs before
and after annealing
Sample Before Annealing After Annealing

Tc1 HC1 TC2 HC2

TPU-70 HS 122.0 ± 0.1 23.0 ± 0.8 123.0 ± 0.1 24.0 ± 0.8

TPU70/GNPM5-NC 131.0 ± 0.2 25.2 ± 0.8 131.0 ± 0.2 24.0 ± 0.2

TPU70/GNPM15-NC 134.0 ± 0.6 25.0 ± 0.5 135.0 ± 0.1 24.0 ± 0.4

TPU70/GNPM25-NC 136.0 ± 0.3 26.0 ± 0.4 135.0 ± 0.4 25.0 ± 0.8

Table 4.7: DSC second heating results for TPU-70 HS and TPU70/GNPs-NCs before
and after annealing
Before Annealing After Annealing
Sample
Tm2 Hm2 Xc2% Tm2 Hm1 Xc2%

TPU-70 HS 172.0 ± 0.2 23.2 ± 0.3 13.0 0.2 173.0± 0.2 23.0 ±0.9 13.0 0.5

TPU70/GNPM5-NCs 182.0 ± 0.6 22.4 ± 0.5 13.0  0.3 184.0± 0.2 21.8± 0.1 13.00.1

TPU70/GNPM15-NCs 183.0 ± 0.5 24.0 ± 0.9 14.0  0.5 184.0 ± 0.1 22.3± 0.4 13.00.2

TPU70/GNPM25-NCs 184.0 ± 0.4 25.0 ± 0.7 15.0  0.4 185.0 ± 0.1 23.5± 1.4 14.40.4

The Tg of soft phase (SP) is not detectable during the first and second heating scan
due to the high amount of HS (70%), leading to restrict the mobility of some of the
SS due to their proximity to HS domain [71]. Conversely, the Tg of HP of neat TPU-
70 HS is recorded in the first heating cycle and before annealing at 47.0 ± 0.5 °C,

152
which shifts to 55.0 ± 0.1, 56.0 ± 0.3 and 58.0  1 °C with the presence of GNPs M5,
M15 and M25, respectively as illustrated in Figure 4.11 A. Taking into consideration
that the DSC technique is not as effective method for measuring Tg as the DMTA
technique. The quantitative values of TgHP before and after annealing with its heat
capacity Cp are summarised in Table A.1 and Table A.2 in Appendix A.

From the improvement in TgHP of TPU-70 HS due to consolidate with GNPs, it can
be speculated that the GNPs form an interfacial interaction with the TPU-70 HS
matrix, causing an increase in the microphase separation of HS and SS [29, 245,
246]. In addition, the interaction with HS can reduce the free volume of the TPU-70
HS molecular chains and inhibits their mobility, leading to an increase in TgHP. The
same results were reported by Razeghi and Pircheraghi [247], who suggested that the
presence of GNPs in a TPU matrix encourages the demixing of SS and HS, leading
to an enhancement in the formation of phase separation. As a result, a
microcrystalline hard domain formed, causing an increase in the Tm of HS. Their
results agreed with those reported by Dan et al. [248], who found that the addition of
nanoclay to a TPU matrix causes a shift in the Tm of HS, as a result of inducing
phase separation.

Furthermore, the annealing treatment causes a minimal increase in Tm of neat TPU-


70 HS and TPU70-NCs and a decrease in Hm in the first heating cycle; in contrast,
in the second heating cycle they are almost unchanged as summarised in Table 4.5
and Table 4.7. It is believed that the reason for this is associated with the samples’
thermal history, which is removed during annealing treatments.

In the cooling cycle and before annealing treatment (see Figure 4.11 C) and Table
4.6), the crystalline temperature (TcHS) for TPU70-NCs increases from 122 C of
neat TPU-70 HS to reach 131°C, 136 °C and 136 C of M5, M15 and M25 GNPs,
respectively, while there is a minimal increase in crystallisation enthalpy (HC) for
all GNP types. It suggests that the GNPs’ role in the TPU-70 NCs acts as a
heterogeneous nucleating agent which accelerates the crystal formation of HS. This
process leads to a rise in its Tc, as demonstrated by other studies [49, 231, 238]. This
result also means an increase in phase separation due to the formation of
microcrystalline hard domain, as mentioned previously in this section. The annealing
treatment does not exhibit any dramatic change in the TcHS and Hc of neat TPU-70

153
HS and its NCs; its value is almost unchanged in comparison to its value before
annealing. Furthermore, depending on the correlation between Hm, the volume
fraction of fillers and the degree of crystallinity, the degree of crystallinity (XC) can
be calculated using the following equation [231]:
∆𝐻𝑚
𝑋𝐶 % = ∆𝐻 × 100 ......... (4.1)
𝑚0 (1−∅)

where:

Hm = heat of fusion of neat TPU-70 HS and its nanocomposites

Hm0 = heat of fusion for 100% crystalline TPU, which is recorded as 172.2 J/g
in accordance with the previous study [249]

 = the nanoparticle fraction in the TPU structure.

From the first and second heating cycles, the XC of neat TPU-70 HS has a moderate
increase for NCs incorporated with GNPM15 and GNPM25. For example, at the
second heating cycle and before annealing, this improvement rises by 8 % and 15.5
% for NCs incorporated with GNPM15 and GNPM25 respectively from that of neat
TPU-70 HS. While, the XC of NCs incorporated with GNPM5 is almost unchanged
in the first and second heating cycle and before and after annealing. This can be
attributed to higher surface area of GNPM15 and GNPM25 which leads to form
more interfacial interaction with cyanate group in hard segment of TPU-70 HS
structure compared with GNPM5 as demonstrated previously in CHNS, XPS and
GPC results (see sections 4.3.1and 4.3.2).

The formations of interfacial interactions with HS make GNPM15 and GNPM25


work as heterogeneous nucleating agent as mentioned earlier in this section, leading
to increase XC of NCs. This behaviour of XC exhibits a good correlation with Hm
according to equation (4.1).

In a similar way, the loading and size of GNPs can reduce the magnitude of XC due
to the decrease in Hm value, which is attributed to poor dispersion or agglomeration
of GNPs within a polymer matrix at high loading. These findings have been
demonstrated in previous studies [208, 239, 250]. In addition, the drop in XC can also
indicate the disruption of the HP orientations by the presence of GNPs, in spite of
their role as nucleating agents [46, 208, 221, 239].

154
4.3.4.3 Dynamic Mechanical Thermal Analysis (DMTA)
Figure 4.12 A and B show the dynamic storage modulus (E’) for neat TPU-70 HS
and TPU70-NCs before and after annealing respectively as a function of temperature.
It can be observed that the whole E’ curves almost exhibit linear steady-state
behaviour below the Tg of SP (TgSP) in the glassy region. Afterwards, the E’ then
starts to decrease gradually and this reduction continues with increasing temperature
due to the thermal expansion of the TPU matrix resulting from the increasing free
volume of molecular chain mobility due to the increase in temperature [241, 251].

A) Non-Annealed B) Annealed
104 4
10

TPU-70HS TPU-70HS
TPU70/5 wt.%GNPM5
Storage Modulus (MPa)

TPU70/5 wt.%GNPM5
TPU70/5 wt.%GNPM15 TPU70/5 wt.%GNPM15
TPU70/5 wt.%GNPM25
TPU70/5 wt.%GNPM25

103 3
10

102 2
10

-100 -50 0 50 100 150 -100 -50 0 50 100 150


Temperature (C) Temperature (C)

Figure 4.12: Log storage modulus, E’, for neat TPU-70 HS and TPU70/GNPs/NCs
based on different GNP sizes (5, 15 and 25m denoted by M5, M15 and M25
respectively) versus temperature (°C): A) before annealing (Non- annealed) and B)
after annealing (Annealed)
This result can be explained by the fact that the TPU-70 HS matrix is rigid below
TgSP, and this rigidity increases with the addition of GNPs. As a result, the value of
E’ increases. Conversely, above this temperature the TPU-70 HS becomes flexible,
leading the value of E’ to decrease with increasing temperature.

In general, the addition of 5 wt.% of GNPs of different sizes (M5, M15 and M25
respectively) to the TPU-70 HS matrix causes an improvement of E’. The extent of
this improvement varies slightly from one type to another. For instance, the M5
GNPs loading possesses the highest value of ’ below TgSP, increasing the E’ for
neat TPU-70 HS from 2717  81 MPa at -100 C to 4846  383 MPa, as opposed to
GNPM15 and GNPM25, which reach 3709  327 MPa and 3754  297 MPa
respectively. Figure 4.13 displays the magnitude of E’ before and after annealing of

155
the neat TPU-70 HS and TPU70/GNPs-NCs at -100, 25 and 100 °C. The quantitative
data of the curve are summarised in the Table A.3 in Appendix A.

Non-Annealed Annealed

-100 C
4
10
Storage Modulus (MPa)

25 C
103
100 C

102

101
PM
5 15 25
S G N NPM NPM
-70H t.% G G
TPU 0/5w 0/5wt.% 0 /5wt.%
7 7 7
TPU TPU TPU

Figure 4.13: Comparison of storage modulus of neat TPU-70 HS and TPU70/GNPs-


NCs based on the different sizes of GNPs, before and after annealing at three
different temperatures (-100 °C, 25 °C, and 100 °C, respectively)

Meanwhile, in the rubbery state or above TgSP, the magnitude of ’ for all types of
TPU70/GNPs-NCs decreases to be very close to each other; however, the results are
still higher than for neat TPU. In the rubbery region (for example, at 25 C), the E’
for TPU-70 HS is 534  16 MPa, which increases upon the incorporation of GNPs to
reach 1362  84 MPa, 1150  17 MPa and 1212  72 MPa for the M5, M15 and M25
GNPs, respectively. These improvements are related to the high modulus of the
GNPs, which enhanced the stiffness of the TPU-70 HS matrix. Besides this
improvement in the stiffness of the matrix polymer, the GNP flakes restricted the
mobility of the SS and HS of the segmented TPU chain. It is believed that this
restriction results from the interfacial interaction between the filler and the TPU
matrix [242]. The same results have been reported in previous studies [16, 241].

The dispersion of GNPs can also affect the mechanical properties of the resultant
NCs. Good dispersion of GNPs within the TPU matrix and strong interfacial
interaction between the filler and polymer matrix can produce NCs with enhanced
156
properties due to their facilitation of the load transfers between them [221, 236, 241,
252]. Nevertheless, the increase in E’ for TPU70/GNPM25-NCs is not as expected in
the glassy region compared with the results recorded for GNPM5 and M15. Halit
[219] ascribes the differences between the expected and experimental E’ values of
polyamide 6/GNPM25 to the folding and rolling of GNPM25 during the
compression-moulding process, due to its large aspect ratio. He proved this
hypothesis using SEM images, which demonstrated that this phenomenon had
occurred. Conversely, the SEM image of TPU70/GNPsM25-NCs in the present study
does not reveal any folding or rolling of GNPM25 in the TPU-70 HS matrix, which
might support the previous hypothesis. However, the E’ magnitude of GNPM25/
NCs is close to that of GNPM15 at high temperatures above TgSP and HP, due to the
increase in mobility of the segmented TPU-70 HS chain.

On the other hand, the annealing treatment of TPU70-NCs (see Figure 4.12 B) has a
negative impact on E’ magnitude, which causes a reduction in its value for all
GNPs/NC types. For example, at 25 °C the E’ for TPU70-NCs drops to be 1080 
186 MPa, 1106  73 MPa and 1157  24 MPa for GNPM5, M15 and M25,
respectively, whereas the value for neat TPU-70 HS increases slightly to 550  7
MPa. This reduction may be related to the microphase-separated disruption in the
TPU-70 HS structure, owing to the restacking of GNP flakes. Typically, during
annealing of neat TPU, the SS and HS are separated to form HP domains and SP
domains, leading to an improvement in microphase separation [47, 82]. Therefore,
the annealing treatment can affect the phase separation of a neat TPU matrix in a way
that can prove the hypothesis of a reduction in hydrogen bonding due to the presence
and formation of a GNP network. Thus, the filler-filler interaction can increase
instead of a filler-polymer interaction, making a reduction in the degree of
microphase separation likely [46, 253, 254]. The Current outcome of E’ for TPU-70
HS and TPU70-NCs after annealing is compatible with the observation that was
reported by Albozahid [46] for TPU-70 HS and TPU70/GNPM15-NCs.

The DMTA technique is considered an accurate technique to determine the Tg of a


polymer, which can be estimated from the highest peak in the tan curves. Figure
4.14 A and B illustrate the tan  curves before and after annealing for neat TPU-70
HS and TPU70/GNPs-NCs. Table 4.8 summarises the quantitative data of TgSP and
TgHP of neat TPU-70 HS and its nanocomposites before and after annealing.

157
A) Non-annealed B) Annealed
0.28 0.28
TPU-70HS
TPU-70HS
0.24 TPU70/5 wt.%GNPM5
0.24 TPU70/5 wt.%GNPM5
TPU70/5 wt.%GNPM15
TPU70/5 wt.%GNPM15
TPU70/5 wt.%GNPM25 TgHP
0.20 0.20 TPU70/5 wt.%GNPM25 TgHP

TgSP
Tan 

0.16 0.16 TgSP


Second
0.12 Relaxation 0.12 Second
Relaxation
0.08 0.08

0.04 0.04

-100 -50 0 50 100 150 -100 -50 0 50 100 150


Temperature (C) Temperature (C)

Figure 4.14: Tan for neat TPU-70 HS and TPU70/GNPs-NCs based on different
GNP sizes (5, 15 and 25 denoted by M5, M15 and M25) A) before annealing and B)
after annealing

Table 4.8: Tan  for neat TPU-70 HS and TPU70/GNPs-NCs based on different
GNP sizes (M5, M15 and M25) before and after annealing

Non- Annealed Non-annealed Annealed


annealed
Sample TgSP TgSP TgHP TgHP

TPU-70 HS -15.3  1.1 -12.4  2.0 67.0  0.9 67.4  0.6

TPU70/5 wt.%GNPM5 -14.7  2.0 -15.0  0.9 77.0  0.6 76.5  1.0

TPU70/5 wt.%GNPM15 -14.8 ± 0.8 -16.0  0.3 78.4  1.6 78.7  2.6

TPU70/5 wt.%GNPM25 -16.0  0.7 -19.0  0.6 78.1  0.7 77.4  1.8

From Figure 4.14 A and B, three different peaks for neat TPU-70 HS and TPU70-
NCs can be observed before and after annealing. The first smaller peak represents the
second relaxation, which has the same value before and after annealing; this is
around -74 °C for TPU-70 HS and approximately the same for NCs. Furthermore,
the second peak represents the Tg for the soft phase (TgSP), which is almost
unchanged by the addition of GNPs before annealing, and changes slightly after
annealing. Finally, the highest peak represents Tg for HP (TgHP), which shifts to a
higher temperature for all GNPs-NCs types before annealing in comparison with neat
TPU-70 HS. The TgHP of TPU-70 HS is detected at 67 C. The TgHP of neat TPU-70
HS increases upon the incorporation of GNPs to be around 77 or 78 C for all GNP
types, as tabulated in Table 4.8.

158
The value of TgHP obtained from the DMTA measurements is greater than its value
obtained from the DSC analysis for all neat TPU-70 HS and its nanocomposites. This
dissimilarity between the two values is associated with the difference in principles of
the two techniques. The same difference between the Tg of SP or HP obtained from
DSC and DMTA methods has been reported in previous studies [47, 222, 231].

This enhancement in TgHP can indicate the role of GNPs in improving stiffness and
load-bearing capacity. Besides, the GNPs confine the mobility of the segmental
molecular chain of the TPU-70 HS matrix and their frictional motion, due to their
interaction with the carbonyl group (C = O) of HP in the TPU-70 HS structure [220,
232, 241, 255]. As demonstrated in sections 4.3.1and 4.3.2 that the present of
GNPM25 and GNPM15 within TPU structure formed a physical interactions with
cyanate and carbonyle group of HS leading to increase oxygene and nitrogyene
amount on the surface of GNPs and then increase the molecular weight of TPU70-
NCs compared to that of neat TPU-70 HS. When the chain mobility is restricted,
relaxation or motion of the chain segments becomes difficult [220, 256, 257]. The
intensity of the tan  peak increases slightly with the addition of GNPs before
annealing and decreases after annealing. This is attributed to the effect of dissipation
energy (E”), which increases at the glass-transition temperature when the chain gains
mobility and could acquire a large amount of dissipated energy during viscous flow.
Conversely, the presence of GNPs prohibits the chain’s molecular mobility, causing
a reduction in dissipation energy as well as a decrease in E’. As a result, the tan 
intensity is slightly increased [258, 259].

Despite these enhancements, the annealing treatment does not show any significant
improvement in the TgSP and TgHP, even though the TgSP shifts to a slightly lower NC
temperature and exhibits a minimial increase for neat TPU. The increase in the TgSP
of neat TPU-70 HS confirms the increase in microcrystal hard domains during
annealing, which restrict the mobility of SS chains and cause an increase in Tg.
Alternatively, a slight decrease in the TgSP of NCs may indicate the GNPs’ role as a
nucleant, resulting in an increase in hydrogen bonds (HB) between HS-HS and a
decrease in HB between HS-SS, reducing the mobility constraints of the SS
molecular chain [82, 247]. A reduction in glass transition has also been observed by
previous studies on graphene/elastomer NC systems; the probable reason was
ascribed to the fillers, which can act as plasticisers. Alternatively, it has been

159
suggested that the dilution effect causes a reduction in the physical entanglements of
the matrix [260]. Another reason for the reduction in the peak intensity of TgHP and
E’ for NCs after annealing could be related to the re-orientation of the filler and an
alteration in their aspect ratio or geometry during annealing. In addition, as tan 
represents the ratio between E” and E’, E” has believed to represent the amorphous
phase within a polymer structure [261]. Therefore, during annealing, the value of loss
factor increases due to the increase in the amorphous phase. This occurs because of
the GNPs’ role in preventing SS and HS from separating and forming HD and SD,
leading to lessen the formation of phase separation.

4.3.5 Melt Rheology of TPU70/GNPs Nanocomposites

The rheological behaviour of the polymer-NCs provides essential information about


their microstructure and processability, as well as their viscoelastic behaviour [20,
93]. The study of the melt rheological behaviour of TPU-70 HS and
TPU70/GNPs/NCs was carried out using oscillatory strain and frequency sweep
measurements.

The oscillatory strain sweep test was conducted to a constant frequency and applied
in order to determine the linear viscoelastic region (LVR). The LVR can be defined
as the region in which the response of the storage modulus (G’) and the loss modulus
(G”) for tested samples are independent of applied strain amplitude [262]. After the
LVR was identified, a frequency sweep was performed at a constant strain within the
LVR (0.2 % in this work) in order to study the dispersion structure of melt
nanocomposites. Both oscillatory strain and frequency sweep measurements were
carried out at 200 C for all samples of neat TPU-70 HS and TPU70/GNPs-NCs.
This temperature was selected based on the processing temperature that was reported
in the melt-compound process for TPU-70 HS and its NCs by Albozahid [46].

4.3.5.1 Oscillatory Strain Sweep


Figure 4.15 shows the data of storage modulus G’ and viscous modulus (loss
modulus G”) for the TPU-70 HS and its nanocomposites containing 5 wt. % of GNPs
(M5, M15 and M25). From this figure, it can be observed that all of the samples
display a Newtonian plateau in the low-strain range, while at high strain amplitude
they exhibit a non-Newtonian plateau.

160
C of TPU70/GNPM25
103 and M15 C of
LVR TPU70/GNPM5 critical strain,  , of
C
neat TPU-70HS
G, G (Pa)

102

G'(TPU-70 HS)
G"(TPU-70 HS)
101 G'(TPU70/5wt.%GNPM5)
G"(TPU70/5wt.%GNPM5)
G'(TPU70/5wt.%GNPM15)
G"(TPU70/5wt.%GNPM15)
G'(TPU70/5wt.%GNPM25)
100 G"(TPU70/5wt.%GNPM25)

100 101 102 103


Oscillation Strain (%)

Figure 4.15: Oscillatory shear measurement results at constant frequency 1Hz for
TPU-70 HS and TPU70/GNPs-NCs at 200 C show the independence of G’ and G”
versus strain amplitude
In this case, the LVR of neat TPU-70 HS continues across a wide range of oscillation
strain (see Figure 4.15 ), which remains constant at G’54 Pa before decreasing at 
= 63% to an order of magnitude lower at the highest limited strain ( = 1500%). The
G” of neat TPU corresponds to a value of 133.8 Pa, which remains constant until  =
10%, at which point it decreases slightly to 80 Pa at the end of the strain sweep ( =
1500%). G” exceeds the G’ value in the neat TPU-70 HS across the whole range of
the strain sweep. When 5 wt. % GNPM5 is added, the G’ of GNPM5/NCs has the
same magnitude as neat TPU-70 HS, except that this starts to decrease at  = 8 %.
Meanwhile, both GNPM15 and GNPM25 cause a significant increase in G’, which
remains constant at 134 Pa and 177 Pa respectively before diminishing at 2 %. This
significant change in LVR with the addition of GNPs indicates a change in the
viscoelastic behaviour of the TPU70-NCs. The point at which the region changes
from a linear to a non-linear viscoelastic regime is known as the critical strain (C)
value [263] as shown in Figure 4.15.

161
The C is significantly altered by the presence of nanoparticles, their size and the
morphology of the nanostructure dispersion of the NCs [264, 265]. The decrease in
magnitude of G’ and G” above C indicates the sudden breakdown of the structure of
the nanocomposites and the breaking up of some important elastic links within the
network of the nanoparticles [262, 264-267]. It is observed that the Newtonian
plateau and C decreases with an increase in the size of the GNPs. Furthermore, it is
clearly seen in Figure 4.15 that the increase in size of the GNPs causes an increase in
G’ and G” values at low strain range or within the LVR. The G” dominates G’ for all
NCs types, especially GNPM25, and at all ranges of the strain sweep, which
indicates the predominantly viscous behaviour of the melted nanocomposites rather
than characteristic elastic behaviour [268].
It is concluded that the presence of GNPs within the TPU-70 HS matrix makes the
resultant NCs very sensitive to low strain amplitude; as a result, the C decreases with
the increase in the size of GNPs as illustrated in Figure 4.15. This might indicate the
ability of GNPs with a larger size (i.e. GNPM15 and GNPM25) to form a fragile
filler networks, leading to a reduction in the LVR region.
From the literature review, Sabzi et al. [25] reported that the addition of two different
sizes of GNPs (larger size NO2 GNPs and small size xGNPs) to poly (lactic acid)
(PLA) by solution mixing could decrease C with increasing volume fraction of both
fillers. They demonstrated that the C could be identified at volume fraction of NO2
GNPs lower than that of xGNP type. They attributed this to the high aspect ratio and
better dispersion of NO2 GNPs. Their results can give a good compatible with the
present study.

4.3.5.2 Frequency Sweep


A frequency sweep was performed at a constant oscillation strain ( = 0.2 %) within
the LVR to study the dispersion of the structure and the linear viscoelastic properties
of TPU-70 HS and its NCs. The frequency sweep also provides further vision into
microstructure changes and network formation of the polymers and their
nanocomposites [269-271].

The shear storage modulus (G’) and loss modulus (G”) versus angular frequency ()
are shown in Figure 4.16 A and B respectively.

162
(A)
105

Storage Modulus, G (Pa)


104

103

102 TPU-70HS
TPU70/5 wt.%GNPM5
TPU70/5 wt.%GNPM15
TPU70/5 wt.%GNPM25
101
100 101 102
(B)
105
Loss Modulus, G (Pa)

104

103

TPU-70HS
TPU70/5 wt.%GNPM5
102 TPU70/5 wt.%GNPM15
TPU70/5 wt.%GNPM25

101
100 101 102
Angular Frequency () (rad/s)

Figure 4.16: Frequency sweep measurements conducted within LVR ( = 0.2 %) at


200C show the values of A) Shear storage modulus, G’ and B) shear loss modulus,
against frequency for TPU-70 HS and TPU70-NCs
It is well known that the rheological properties are very related to the dispersion of
polymer/nanocomposites and the G” is less sensitive with the filler incorporation
than the value of G’ and complex viscosity, * [103, 105]. However, from Figure
4.16 A and B, it can be observed that the G’ and G” for TPU70-NCs increases
dramatically with the incorporation of GNPs (M5, M15 and M25) compared to that
of neat TPU-70 HS. This improvement is attributed to the ability of GNPs as known

163
fillers with high surface area to increase the interfacial interaction with TPU-70 HS
matrix and prohibit the chain mobility at melt state. It can also be observed from this
figure that all of the samples of NCs have roughly similar melt rheological behaviour
across various GNP sizes.

Furthermore, the G’ and G” of all samples increase with increased angular


frequency, and the highest increase ascribed to the NCs incorporated with GNPM25.
At low frequency (less than 10 rad/s), the G’ and G” for NCs, especially that one
incorporated with GNPM25 have a plateau nearly to be stable and less dependent on
, while at high frequency their values become highly dependent on the frequency.
This behaviour means that the TPU70/nanocomposites at a low frequency nearly
approached its rheological percolative network threshold. The rheological
percolation threshold occurs when the G’ of nanocomposites becomes independent
of angular frequency at low-frequency ranges and dependent at high thereof. In this
state, the rheological behaviour of NCs starts to move from liquid-like behaviour to
solid-like behaviour. Therefore, these results indicate that the GNPs especially
GNPM25 start to form rheological percolated network at this concentration (5 wt. %)
within TPU-70 HS matrix, changing to solid-like behaviour [100, 101, 103, 105].
Meanwhile, at high frequency, this behaviour transforms due to break down elastic
structure network of GNPs [25, 268].

The data of complex viscosity (*) of TPU-70 HS and TPU70-NCs is shown in


Figure 4.17, in which the complex viscosity (*) of GNPs (M5, M15 and M25) at 1
rad/s increases to two or three times its value for neat TPU-70 HS. All samples of
GNPs/NCs show the same trend of *, which decreases with increasing angular
frequency. A rapid decrease occurs when  is greater than 10 rad/s. Both GNPM15
and M25 have fewer drops in *, when  is more than 50 rad/s, in contrast with
GNPM5. The * for GNPM5 continues to drop, reaching a value lower than that for
neat TPU-70 HS when  is above 200 rad/s [104, 272, 273]. The behaviour of
NCs/GNPM5 might be ascribed to the small size of GNPM5, which leads to less
interfacial interaction with TPU chain as demonstrated previously (see sections
4.3.4.3 and 4.3.4.1). Therefore the presence of GNPM5 in the entangled and confined
chains can lead to increase their melt free volume causing a high drop in viscosity
[104, 105, 274]. In general, the drop in viscosity results from the robust shear-

164
thinning behaviour of the NCs and neat TPU-70 HS, comparable to a molten state,
and is also due to the drag of the polymer mobility at high frequency [268, 275].

104
TPU-70HS
Complex Viscosity (, Pa.s)

TPU70/5 wt.%GNPM5
TPU70/5 wt.%GNPM15
TPU70/5 wt.%GNPM25
103

102

101
100 101 102
Angular Frequency () (rad/s)

Figure 4.17: Complex viscosity, *, of neat TPU-70 HS and Its NCs frequency sweep
measurements conducted within LVR ( = 0.2%) show the influence of incorporating
with different sizes of GNPs at same loading 5 wt. %.

To summarise, the magnitude of * for TPU70/GNPs/NCs surpasses the value of


neat TPU-70 HS despite their drop at high , which remains higher for the entire
range of  except GNPM5. This result indicates that the GNPs, and GNPM25 and
M15 in particular, increase the resistance of polymer chain flow and deformation in a
molten state. This outcome can also be attributed to the strong interfacial interaction
with GNPs, and especially GNPM25, which possesses the highest aspect ratio.

From the literature review, Kim and Macosko [106] demonstrated that incorporated
G within PC matrix leading to reach rheological percolation threshold between 3 and
5 wt.% of G, while PC matrix incorporated with functionalised graphite sheets (FGS)
become rheological percolated at wt.% between 1 and 1.5. They attributed this
discrepancy in rheological percolated to the improvement in dispersion and
interaction of FGS with PC matrix owing to functionalised G.

165
Poteschke et al. [105] study the rheological behaviour of CNTs consolidated with PC
matrix by using various concentration arranging from 0.5 to 15 wt.% of CNTs. The
nanocomposites materials were prepared using a melt compounding process and then
compression moulding technique. They found that the viscosity and elastic melt
properties denoted by G’of NCs achieved a Newtonian plateau at low frequency at 2
wt. % CNTs. They considered this percentage as a rheological percolation threshold,
and by comparing with other studies, they demonstrated that this weight percentage
is lower than other PC/NCs consolidated with black carbon or carbon nanofibers as a
filler. The researchers attributed this lower percolation threshold to the high aspect
ratio of CNTs compared with carbon black and carbon nanofibers.

4.3.6 Mechanical Properties

The development of inorganic nanofillers as reinforcements in a polymer matrix has


been proven as a successful way to produce a nanocomposite material with excellent
properties. Not only can the size and geometry of the fillers be considered a crucial
factor for improving the physicochemical properties of a matrix, but their dispersion
is also a significant feature which controls the properties of the resultant NCs. At
high loading, agglomeration of the fillers in stacks or bundles can lead to premature
failure, as this feature can act as a defect or stress-concentration point within the
matrix. As such, this section focuses on two mechanical testing approaches (tensile
test and flexural test) which were carried out during this investigation.

4.3.6.1 Tensile Test of TPU-70 HS and TPU70/GNPs-NCs


A tensile test is one of the most straightforward and common methods of mechanical
testing. It is used widely to investigate the influence of inorganic fillers on the
behaviour of a polymer matrix under static loading. The results of this test are plotted
as a stress-strain curve; from this curve, all of the tensile properties can be calculated,
including elastic modulus (E), yield strength (y - the maximum stress before plastic
deformation), tensile strength (u), and elongation at break (u).

In this study, a tensile test was carried out to scrutinise the effect of loading GNPs
and their different sizes on the tensile properties of neat TPU-70 HS and
TPU70/GNPs-NCs. Representations of stress-strain curves for neat TPU-70 HS and
its nanocomposites before and after annealing are shown in Figure 4.18 A and C,

166
respectively. Figure 4.18 B and D represents a zoomed-out view of the elastic
region of the initial part of the tensile stress-strain curve.

A) Non-annealed B) Non-Annealed
40 30
TPU-70HS
35 TPU70/5wt. %GNPM5
TPU70/5wt.%GNPM15 25
30 TPU70/5wt. %GNPM25
Stress (MPa)

20
25

20 15

15
10
TPU-70HS
10 TPU70/5wt. %GNPM5
5 TPU70/5wt.%GNPM15
5 TPU70/5wt. %GNPM25

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 0.1 0.2 0.3
C) Annealed D) Annealed
40 30
TPU-70HS
35 TPU70/5wt. %GNPM5
TPU70/5wt.%GNPM15 25
30 TPU70/5wt. %GNPM25
Stress (MPa)

20
25

20 15

15
10 TPU-70HS
10 TPU70/5wt. %GNPM5
5 TPU70/5wt.%GNPM15
5 TPU70/5wt. %GNPM25
0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.05 0.10 0.15 0.20 0.25 0.30
Strain (mm/mm) Strain (mm/mm)

Figure 4.18: Stress-strain curves for TPU-70 HS and TPU70/GNPs-NCs: A, B)


before annealing and C, D) after annealing. The right hand curves B and D
represent a zoomed-out view of the elastic region of the initial part of the tensile
stress-strain curve A and C respectively
It can be seen that the neat TPU-70 HS and TPU70-NCs with GNPM5 do not display
a clear yield point before annealing; the same is true for all test samples after
annealing. As a result, the 0.2% offset method was used in order to estimate the
magnitude of y. It can be clearly seen that, before and after annealing, the tensile
stress of neat TPU-70 HS increases steadily with increasing strain until reaching
breaking point; the same behaviour is observed for NCs based on GNPM5. On the
other hand, TPU70/GNPM15 and M25 samples only exhibit a very clear yield point
before annealing; their tensile stress starts to decrease after this point and then
increase gradually until the point of fracture. In contrast, after annealing, their trend
behaviour becomes similar to that of the neat TPU-70 HS and GNPM5/NCs samples.
The behaviour of the GNPM15 and M25/NC samples is compatible with their
behaviour recorded in the WAXD results. Conversely, both GNPM15 and M25 have

167
the same curve trend before annealing, but the trends of both altered after annealing,
reducing the peak intensity. This outcome is due to the increase in the amorphous
phase and the reduced crystalline phase (ordering of HS; see Section 4.3.3.3).

Overall, the 5 wt.% loading of GNPs based on different aspect ratios causes a
significant increase in tensile modulus and yield strength before annealing and a
marginal decrease in elongation at break. There is also a slight decrease in tensile
strength. Figure 4.19 presents the results of the tensile properties before and after
annealing of pristine TPU-70 HS and TPU70/GNPs-NCs. Their corresponding
values are tabulated in Table A.4 in Appendix A.

1.0
(A) (B)
28
Modulus of Elasticity, E (GPa)

0.8 Yeild stress, y,(MPa) 24 Non-annealed


Non-annealed
Annealed Annealed
20
0.6
16

0.4 12

8
0.2
4

0.0 0
5 15 25 5 15 25
NPM GN
PM
GN
PM S NPM GN
PM
GN
PM
0HS %G 70H %G
U-7 wt wt% wt% U- wt wt% wt%
TP 70/5 70/5 70/5 TP 70/5 70/5 70/5
TPU TPU TPU TPU TPU TPU

(C) (D)
40
400
Non-annealed Non-annealed
Elongation at break (mm/mm) %

35
Annealed 350 Annealed)
Tensile Strength (MPa)

30
300
25
250
20
200

15 150

10 100

5 50

0 0
5 15 25 PM
5 15 25
NPM PM PM GN PM PM
S GN GN -70H
S GN GN
U-70H wt%G wt% wt% P U /5wt%
/5wt%
/5wt%
TP 70/5 70/5 70/5
T
TPU
70
TPU
70
TPU
70
TPU TPU TPU

Figure 4.19: Tensile properties of neat TPU-70 HS and TPU70/GNPs-NCs based on


different sizes of GNPs before and after annealing
For example, the tensile modulus (E) of pristine TPU-70 HS is 0.20 GPa, which
increases to reach 0.28 GPa, 0.55 GPa, and 0.74 GPa of NCs for GNPM5, M15 and
M25, respectively as shown in Figure 4.19 A. These results indicate that GNPM25
enhances the tensile modulus of TPU matrix approximately fourfold, as opposed to
the twofold improvement offered by GNPM5. This is because the high aspect ratio of

168
GNPM25 leads it to possess higher interfacial interaction with the TPU-70 HS
matrix than GNPM5. As a result, there is a good chance that stress would be
transferred from the polymer matrix to the dispersed phase (GNP flakes) [107, 219,
241]. Moreover, the good interaction between GNPs and pristine TPU causes the
GNPs to function as a skeleton in the matrix, which reduces the mobility of the
polymer molecular chain and improves the stiffness of the resultant NCs [232, 241,
276, 277].

Yield strength, y (see Figure 4.19 B), follows the same trend as tensile modulus
before and after annealing, in that it increases with increasing GNP aspect ratio. It
can be observed that the y of NCs rises to 12.0 MPa, 19.4 MPa, and 27.5 MPa when
incorporated with GNPM5, M15, and M25, respectively, in comparison with 9.8
MPa of pristine TPU-70 HS. In contrast to these results, after annealing both E and
y decrease when the NCs are incorporated with GNPM15 and M25, while there is
an improvement in neat TPU-70 HS and GNPM5/NCs. It is believed that the
physicochemical interaction between the HS of the TPU structure and the GNPM15
and M25 flakes confines the movement of HS during annealing, resulting in a
reduction in the microcrystalline hard domain and then a decrease in phase
separation. This hypothesis is compatible with the suggestion made by Laity et al.
[80], who reported that the decrease in the mechanical properties of TPU can be
attributed to the reduction in phase separation of HS and SS by increased demixing
between them. On the contrary, due to its small surface area and less physical
interaction with TPU-70 HS matrix, GNPM5 can act as a heterogeneous nucleate
agent during annealing treatment, leading to an enhancement in the microcrystalline
hard domain and enhanced phase separation. For the pristine TPU-70 HS, the
annealing treatment helps to increase the ordering of HS and improve micro-phase
separation. Thus, the degree of phase separation increases and an improvement in
mechanical properties occurs.

Besides an improvement in E and y of the pristine TPU and TPU70-NCs before


annealing, the tensile strength u and elongation at break decrease, as illustrated in
Figure 4.19 C and D. Compared with pristine TPU-70 HS, the tensile strength for
NCs drop from 26 MPa of neat TPU-70 HS to 20.5 MPa and 23.5 MPa for GNPM5
and M15 respectively, while GNPM25 is almost unchanged. Meantime, the
elongation at break of NCs reduced from 299% of pristine TPU-70 HS to 172%,

169
223% and 194% for GNPM5, M15 and M25 respectively. This result can be ascribed
to the role of GNPs, which seem to increase filler-filler interaction in comparison
with filler-matrix interaction, leading at some point to aggregation and then
prohibiting strain hardening, as reported by the previous studies [29, 231, 278]. In
addition, when pristine TPU is subjected to tensile loading, its molecular chain tries
to orient itself in the direction of the tensile elongation, leading to maximum
intermolecular interaction [107, 279]. Therefore, it is suggested that the presence of
nanofillers interrupt the molecular arrangements, causing this reduction in tensile
strength and elongation.

The same tensile results for NCs based on various nanofiller sizes were reported by
previous studies. For example, Halit [219] revealed that GNPM25 yields greater
improvement of tensile properties for polyamide 6/NCs than GNPM5 due to its
higher surface area. In addition and from the literature review, Choi et al. [107]
found that the tensile modulus of the TPU matrix is reinforced by two different sizes
of graphene sheet, improving more with the graphene sheet with a larger surface area
than the smaller one. It was then concluded that the stress transfer by the higher
surface area of the graphene sheet is more efficient than the smaller one due to the
increase in interfacial interaction with the matrix. Furthermore, other studies have
focused on studying nanofiller dispersion and their uniform distribution in different
polymer structures, and their subsequent effect on tensile properties [26, 238, 280,
281]. For instance, Suh and Bae [26] studied the effect of GNP dispersion on the
mechanical properties of PTFE. They concluded that the yield strength of PTFE
reinforced by 3 vol.% GNPs increased by 60% in comparison with pristine PTFE.
They explained this enhancement by the effect of GNPs on preventing chain
movements due to their random orientation.

The annealing treatment caused an extra reduction in elongation for NC samples and
a reduction in TPU-70 HS. Moreover, there is a slight reduction in tensile strength
for GNPM25-NCs and a moderate increase in pristine TPU-70 HS, GNPM15, and
GNPM5-NCs. This result can be attributed to the effect of GNPs on reducing
microphase separation due to their interaction with HS, as mentioned previously [46,
282]. The same results for TPU70/GNPM15 before and after annealing were
reported by Albozahid [46].

170
4.3.6.2 Flexural Properties of TPU-70 HS and TPU70-NCs
In this study, a flexural test was carried out in order to investigate the effect of
various sizes of GNPs on the flexural properties of neat TPU-70 HS and TPU70-
NCs. Figure 4.20 shows the typical stress-strain curve for the flexural test of the neat
TPU-70 HS and its NCs before and after annealing.

A) Non-Annealed B) Annealed
60 60
Neat TPU-70HS Neat TPU-70HS
TPU70/5wt%GNPM5 TPU70/5wt%GNPM5
50 50 TPU70/5wt%GNPM15
TPU70/5wt%GNPM15
Flexural Stress ( MPa)

TPU70/5wt%GNPM25 TPU70/5wt%GNPM25
40 40

30 30

20 20

10 10

0 0
0.00 0.02 0.04 0.06 0.08 0.10 0.00 0.02 0.04 0.06 0.08 0.10
Flexural strain ( mm/mm) Flexural strain ( mm/mm)

Figure 4.20: Typical flexural stress-strain curve for neat TPU-70 HS and
TPU70/GNPs-NCs based on various size of GNPs (5,15 and 25 m denoted by M5,
M15 and M25 respectively): A) before annealing (Non-annealed) and B) after
annealing (Annealed)
As a function of the GNP size, all of the NCs exhibit the same trend of the flexural
stress-strain curve, which starts to go gradually upwards until reaching the maximum
peak stress then fracturing immediately, while neat TPU-70 HS samples do not break
any more. Figure 4.21 shows the type of flexural failure that happened to neat TPU-
70 HS and NCs samples. From this figure, it can be clearly observed that TPU-70 HS
samples do not break whilst all the NCs samples subjected to brittle fracture and
broke immediately.

Neat TPU-70 HS

TPU70/5 wt.%GNPM5-NCs

TPU70/5 wt.%GNPM15-NCs

TPU70/5 wt.%GNPM25-NCs

Figure 4.21: Typical sample of Neat TPU-70 HS and TPU70/GNPs-NCs after


subjecting to flexural test shows a type of flexural failure
171
The flexural stress-strain curve of neat TPU-70 HS samples starts to increase
gradually until reaching peak stress; this then also decreases gradually without
breaking the sample as shown in Figure 4.21. In general, all the NC samples show a
reduction in flexural strain failure before and after annealing; however, their stress-
strain curve trends are higher than those of the neat TPU-70 HS matrix, pointing to
the enhancement in flexural modulus and strength. Contrary to expectations, the
GNPM25/NCs exhibit lower failure strain in both cases in comparison to GNPM5
and M15. Figure 4.22 A, B, and C shows the flexural strength (FS), flexural modulus
(FM) and flexural failure strain (FFS) respectively for neat TPU-70 HS and TPU70/
NCs. The corresponding values are listed in Table A.5 in Appendix A.

(A) (B)
50 1.8
Non-Annealed
Non-Annealed
Annealed 1.6
45 Annealed
Flexural Strength (MPa)

Flexural Modulus (GPa)


1.4
40 1.2

1.0
35
0.8
30 0.6

0.4
25
0.2
20 0.0
5 M15 M25 5 15 25
HS NPM GNP GNP HS NPM NPM NPM
U-70 %G wt % wt % U-70 %G %G %G
eat TP 70 /5wt 0 /5 0 /5 ea t TP 70 /5wt 70 /5wt 70 /5wt
N TPU 7 7 N TPU
TPU TPU TPU TPU

12
(C)
Non-annealed
Annealed
10
Flexural Failure strain (%)

0
5 15 25
0HS NPM NPM NPM
TP U-7 5wt%G wt %G wt %G
Nea
t 70/ 70/5 70/5
TPU TPU TPU

Figure 4.22: Flexural properties of neat TPU-70 HS and TPU70/GNPs-NCs based


on various size of GNPs (5,15 and 25 m denoted by M5, M15 and M25
respectively) before and after annealing
From Figure 4.22 B the flexural modulus of neat TPU-70 HS is 0.7 GPa, which rises
with the presence of GNPs to reach 1.1 GPa, 1.2 GPa and 1.2 GPa for NCs filled
with GNPM5, M15, and M25, respectively. This improvement can be attributed to
the improvement in stiffness ascribed to the high modulus of GNPs [106, 283].
Furthermore, the FS (see Figure 4.22 A) is increased by the addition of GNPs to

172
reach 39 MPa, 43 MPa, and 42 MPa for TPU70/GNPM5, M15, and M25,
respectively, in comparison with 37 MPa for neat TPU-70 HS. It is clear that, for
flexural modulus, there is no significant disparity in its values for the various
GNPs/NCs. However, for FS, the GNPM15/NCs and M25 have the higher values,
which is compatible with expectations due to their large aspect ratio and good
interfacial adhesion with the TPU-70 HS matrix. From the literature review, Onyu et
al. [23] investigated the influence of two different GNP sizes (GNP-M-5 had a 5
m-average diameter and GNP-H-100 had an average diameter of 200µm) on the
electrical and mechanical performance of GNPs/PP/NCs. They revealed that the
GNP-H-100 gave better flexural performance than the smaller GNP-M-5 due to its
larger surface area and good interaction with the PP matrix.

The disparity in the results from different samples is not only related to the GNP
size; the dispersion of nanofillers can be considered a crucial factor that influences
the NCs’ performance. Wang et al. [215] studied the effect of GNP size and their
dispersion on the mechanical properties of an epoxy matrix by dispersing two
different sizes of GNPs (GNP-5 measuring 5 µm and GNPC750 measuring less than
1 m) in the matrix. They concluded that the addition of the larger GNPs (5 m) to
the epoxy matrix causes a significant improvement in flexural and tensile modulus
due to its high aspect ratio and uniform distribution, which limits the chain mobility
of the epoxy matrix. However, the strength of the resultant GNP-5/epoxy NCs was
reduced due to their weak interfacial interaction with the matrix. Therefore, it was
concluded that the uniform dispersion may affect the flexural performance as well as
the GNPs’ size. Kim and Macosko [106] reported that the incorporation of GNP
fillers in a polycarbonate (PC) matrix gave excellent improvement in stiffness if they
were dispersed in a homogeneous way. Their results were consistent with other
studies, which revealed that the stiffness of the resultant NCs was improved with
large particle size; however, this observation was true on the condition that
homogeneous distribution was achieved, as the opposite case can reduce the
interfacial interaction, leading to nanofiller agglomeration [284-287].

It can be observed from Figure 4.22 A, B and C that the annealing treatment has a
particularly dramatic effect on FFS for TPU70/GNPM15 and M25 among the other
flexural properties and in comparison with their values before annealing. While the
GNPM25/NCs exhibit a reduction in all of their flexural properties (FS, FM and

173
FFS) after annealing, the other samples, especially neat TPU-70 HS and
GNPM5/NCs, exhibit a slight improvement in all of them. This result can be
explained by the fact that during annealing treatment of the neat TPU-70 HS there is
sufficient time for HS to reorder, leading to increase the formation of the
microcrystalline hard domain and improved microphase separation. These results are
consistent with tensile-test results, where the same reason can be applied.

For the NC results, the GNPs can act in two ways. Firstly, they can prevent chain
mobility and lead to a reduction in the degree of crystallinity. This then causes the
stiffness performance of NCs to deteriorate, in the same way, observed for GNPM25.
Secondly, they can act as a heterogeneous agent, causing an improvement in the
degree of crystallinity; in this case, the NCs’ stiffness is improved, as is the case for
GNPM5. As observed from the tensile, CHNS, XPS and GPC results (see sections
4.3.1 and 4.3.2), GNPM25 possesses a good interfacial interaction with HS in the
TPU matrix, due to the fact that the annealing treatment has a more negative impact
on its mechanical properties than for other sizes. That is, during this treatment the
GNPM25 prohibits the movement of HS to improve its ordering, leading to a
reduction in the microphase separation of the TPU structure and reducing its
properties.

4.3.7 Testing Conductive Properties

One of the unique properties of GNPs is their high electrical and thermal
conductivity, which makes GNPs the most promising filler for producing polymer
nanocomposite with outstanding improvement in their intrinsic conductive
properties. However, the methods for dispersing and distributing GNP flakes in the
polymer matrix can directly affect the enhancement in electrical and thermal
conductivity of the resultant NCs. Therefore, it is vital for GNPs to form a
conductive path that is sufficient for transporting electrons in the polymer NCs. This
section focuses on studying the effect of GNP size on the electrical and thermal
conductivity of TPU-70 HS and TPU70/GNPs/NCs.

4.3.7.1 Electrical Conductivity


In the present work, the electrical conductivity for TPU-70 HS and TPU70/GNPs-
NCs was calculated in two directions, in-plane (longitudinal) and out-of-plane
(transverse). This was done in order to investigate the effect of GNP size on the

174
electrical conductivity for neat TPU and TPU70-NCs, as well as the effect of the
samples’ dimensions and GNP distributions. All samples for electrical conductivity
measurements were prepared using compression moulding. The electrical
conductivity test of all samples was carried out by applying AC current using a two–
probe impedance spectrometer in order to measure the impedance for these samples.

The variation of logged electrical conductivity versus log frequency in the range of 1
to 10^6 Hz for neat TPU-70 HS and TPU70/GNPs-NCs is illustrated in Figure 4.23
A and B. The TPU70/GNPM5-NCs show electrical conductivity that is dependent on
frequency, which indicates that these NCs are still insulated in both directions (in-
plane and out-of-plane). It also reveals that the percolation threshold of GNPM5 is
not achieved at 5 wt.% due to its small size (see frequency sweep in section 4.3.5.2).

(A) (B)
2 TPU-70HS 2 TPU-70HS
10 10
TPU70/5wt.%GNPM5 TPU70/5wt.%GNPM5
Electrical Conductivity (S/m)

TPU70/5wt.%GNPM15 TPU70/5wt.%GNPM15
100 In-plane direction 100 TPU70/5wt.%GNPM25
TPU70/5wt.%GNPM25

10-2 10-2
Out-of-Plane direction

10-4 10-4

10-6 10-6

10-8 10-8

10-10 10-10
100 101 102 103 104 105 106 100 101 102 103 104 105 106
Frequency (Hz) Frequency (Hz)

Figure 4.23: Electrical conductivity versus frequency for TPU-70 HS and


TPU70/GNPs-NCs in two directions: A) In-plane direction and B) Out-of-plane
direction
The percolation threshold can be defined as the ability of the filler to form a
conductive network, which leads to a sudden increase in the electrical conductivity of
the composites [288]. Meanwhile, the NCs with GNPM15 and GNPM25 show
independent electrical conductivity at frequencies below 1500 Hz for the in-plane
direction (Figure 4.23 A) and 1000 Hz in the out-of-plane direction (Figure 4.23 B).
This result indicates that both GNPM15 and M25 reach their percolation threshold in
each direction at 5 wt. % due to their high specific area and/or aspect ratio.

When the frequency value is high, the electrical conductivity for GNPM15 and M25
becomes dependent on frequency, meaning that it increases with frequency. These
results are compatible with those reported by Albozahid [46], who reported that the

175
TPU70/GNPM15-NCs prepared by in-situ polymerisation became conductive
material at 5 wt.% of GNPs. Those results of Albozahid [46] confirmed those of
Yousefi et al. [289], who indicated the same trend of TPU nanocomposites using GO
and rGO nanofillers. Ramos et al. [290] used CNT and carbon black (CB) as fillers
with TPU, and inferred that CNT reaches the percolation threshold in a low
concentration compared to CB due to the high aspect ratio of the former. In general,
for NCs reinforced by filler with high aspect ratio, their percolation threshold was
observed at low concentration, as reported by previous studies [288, 291, 292].

As seen in Figure 4.23 A and B, the magnitude of electrical conductivity of


GNPM15 and M25 in the in-plane direction is around ten times higher than its value
in the out-of-plane direction. This result demonstrates that the sample dimensions
and the GNP distribution affect the formation of conductive network by nanofillers,
as illustrated in Table 4.9. In addition, the formation of conductive network is
strongly influenced by particle orientation, which plays a crucial role. For instance, a
particle with parallel alignment needs a greater amount to reach the percolation
threshold, as concluded by previous studies [293, 294].

Table 4.9: Electrical conductivity of TPU-70 HS and TPU70/GNPs-NCs

Electrical conductivity,  Electrical conductivity, 


(S/m) (S/m)
Samples
In-plane (longitudinal) Out-of-plane (transverse)
direction direction
Neat TPU-70 HS no no
TPU70/5wt. %GNPM5-NCs no no
TPU70/5wt.% GNPM15-NCs 1.14  10-3  0.54  10-3 7.1  10-5  1.7  10-5
TPU70/5wt.%GNPM25-NCs 1.16  10-4  0.45  10-4 1.0  10-5  0.4  10-5

4.3.7.2 Thermal Conductivity


A crucial factor which governs the thermal conductivity of a material is the phonon
or lattice vibration. As a result of this factor, the formation of a conductive thermal
network is not only enough to cause a significant enhancement in the thermal
conductivity of the NCs, but it also necessitates a strong filler-polymer interface.
Figure 4.24 shows the effect of GNP size on the thermal conductivity of neat TPU-70
HS and TPU70-NCs for the same loading of GNPs.

176
0.9

0.8
20 C

Thermal conductivity (W/ m.K)


30 C
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0.0
5 15 25
H S GNPM G NPM GNPM
-70 wt.% /5 wt.% wt.%
TPU U70/5 70 70/5
TP T P U TP U

Figure 4.24: Thermal conductivity of neat TPU-70 HS compared to TPU70/GNPs-


NCs based on the various size of GNPs
In this figure, the results obtained emphasise that the thermal conductivity of the neat
TPU-70 HS and its nanocomposite increase with increasing temperature and aspect
ratio. For example, at 20 C, the thermal conductivity for neat TPU-70 HS increases
from 0.35 W/m.K to 0.56 W/m.K, 0.69 W/m.K, and 0.75 W/m.K for NCs
incorporated with GNPM5, M15, and M25, respectively. The same trend is observed
at T = 30 C. The greatest improvement is caused by GNPM25, which increases the
thermal conductivity of neat TPU by 116 %, whereas GNPM5 and M15 show
improvements of around 60 % and 96 %, respectively, substantially lower than the
improvement of GNPM25. Although the thermal conductivity for neat TPU is
increased by the addition of GNPs, it is not as dramatic as expected, considering the
higher thermal conductivity of GNPs (3000 W/m.K as tabulated in Table 3.6, section
3.2.2.1). This finding implies that the contact between GNPs and TPU-70 HS is
rather weak, and that this produces interfacial thermal resistance at the interface
between GNPs and TPU-70 HS, which obstacles heat flow [295].

Overall, GNPM25 produced the highest improvement in thermal conductivity of


TPU70-NCs in comparison with GNPM15 and M25. This outcome is due to
differences in aspect ratio of GNP types; GNPM25 possesses a higher aspect ratio,
meaning that they have more flake boundaries and edges, which cause phonon
scattering [296]. Thus, the increase in phonon scattering causes a decrease in
interfacial thermal resistance, and the consequence is an improvement in thermal

177
conductivity. These thermal conductivity results are compatible with those of Halit
[219], who reported that PA6/GNPM25 had higher thermal conductivity than
PA6/GNPM5. A similar result was reported by Kim et al. [109], who was mentioned
in the literature review, when they prepared a nanocomposite based on polycarbonate
reinforced by five GNP types of varied sizes and thicknesses [109]. Their results
demonstrated that GNPs with large size and thickness caused an effective
enhancement in thermal conductivity and composite ability for heat dissipation. The
same results were discovered in previous studies on the effect of filler size and
dimension on the thermal conductivity of NCs [297, 298], which substantiate the
current results.

4.4 Summary

• In this study, three different GNP sizes, grade M, with average particle diameters
of 5 m, 15 µm and 25 m have been used to reinforce a TPU-70 HS matrix using
in-situ polymerisation. The effect of sonication on the GNP flakes was
characterised using TEM, SEM and WAXD techniques. These techniques proved
that there was no change in the shape of the GNP flakes due to the short time and
low frequency used throughout the sonication process. In addition, WAXD results
confirmed no change in the location of GNP peaks, with both peaks still in the
same place (26.6 and 55°, respectively), indicating no alteration in their
interlayer spacing.
• To study the dispersion of GNP flakes in the TPU-70 HS matrix, SEM and OM
techniques were used. Both techniques confirmed that GNPs had good dispersion
in TPU-70 HS. It was concluded from the WAXD results that the GNP flakes
maintained their peaks inside the TPU-70 HS matrix, which indicated that there
was no alteration in their crystallography plane.
• The influence of both GNP sizes and annealing treatment on the thermal and
mechanical properties of the neat TPU-70 HS and its composites were
investigated. The TGA results indicated that all GNP flakes underwent
improvement in Tdonset of TPU-70 HS, Tdmax1 for HS and Tdmax2 for SS.
• The DSC results indicated an increase in TC and Tm of TPU-70 HS upon
incorporation with GNPs and moderate improvement in XC1 % and XC2 % for
NCs. Conversely, the annealing treatment did not appear to have any dramatic

178
effects on TC, Tm1, Tm2, XC1, XC2, Hm1, or Hm2. In addition, the DMTA results
showed that the storage modulus improved in the glassy region below the TgSP and
then decreased gradually above this temperature for all NC samples. This result
confirmed that the stiffness of neat TPU-70 HS was improved due to the presence
of GNP flakes, although it decreased in the rubbery state but E’ was still greater
than that for the TPU-70 HS matrix. The tan  results showed an improvement in
TgHP by the addition of GNP flakes in comparison with TPU-70 HS. Furthermore,
the annealing treatment caused a slight reduction in E’, TgSP and TgHP, which
indicated the effect of GNP flakes on reducing the degree of hydrogen bonds due
to their interfacial interaction with HS and the subsequent reduction of the degree
of microphase separation of the TPU-70 HS matrix.
• The investigation of the effect of GNP flake size on rheology properties of neat
TPU-70 HS and its nanocomposites was accomplished using the oscillation
amplitude and oscillation frequency. It was concluded that the GNPM25 had a
greater improvement in G’ and * due to its high surface area, aspect ratio, and
good dispersion compared with GNPM5 and GNPM15.
• The tensile and flexural tests showed an improvement in tensile and flexural
modulus before annealing. These results were attributed to the enhancement in the
stiffness of the neat TPU-70 HS matrix due to the incorporation of GNPs, which
had a higher modulus. Finally, the annealing treatment caused a slight reduction in
tensile and flexural modulus for GNPM25 and GNPM15-NCs, while there was an
improvement in their magnitude for neat TPU-70 HS and TPU70/GNPM5-NCs.
• The electrical conductivity test revealed that the NCs started to be conductive at 5
weight percentiles of the GNPM15 and M25, which was considered their
percolation threshold, while GNPM5 needed a greater amount to reach this step
due to its smaller size. The thermal conductivity test reported that the
TPU70/GNPM25/NCs had a higher magnitude than other types.

Overall, the presence of GNP flake within TPU-70 HS caused a significant


improvement in its electrical, thermal, and mechanical properties; more specifically,
GNPM25-NCs yielded the best performance. For instance, the incorporation of
GNPM25 within TPU-70 HS matrix caused an improvement in thermal conductivity,
tensile modulus and flexural modulus by 116 %, 270 % and 71% respectively. These

179
properties made the TPU70/GNPs-NCs suitable for use in different applications such
as electronic, packaging and automotive functions.

The annealing treatment showed an improvement in mechanical and thermal


properties of neat TPU-70 HS and moderate improvement in NCs incorporated with
GNPM5 while NCs incorporated with GNPs M15 and GNPM25, in particular,
showed a decrease in most thermal and mechanical properties. Therefore, all of these
matrices were used to form multiscale composites in order to study the effect of
GNPs size and annealing treatment on TPU70/CF composites in terms of their
electrical, thermal and mechanical properties. The best MSCs were then found, of
which different percentages (0.5 wt. %, 2.5 wt. % and 5 wt. %) can be used in order
to study the effect of GNPs on its damage tolerance and interlaminar fracture
toughness properties.

180
Chapter 5: Results and Discussion of
TPU70/GNPs/CF Multiscale Composites (MSCs)

5.1 Introduction

Carbon fibre-reinforced thermoplastic composites have attracted great attention in


recent decades as alternative materials to carbon fibre/thermoset composites.
Thermoplastic polyurethane is selected in this study to be manufactured with carbon
fibre, due to its individual unique segmented structure and properties. TPUs have an
excellent resilience and toughness, as well as good abrasion resistance and due to its
adhesion properties which can negate the surface treatments of carbon fibre and
circumvent their weaknesses. In general, in the field of fibre-reinforced polymers, the
fibre is responsible for the in-plane properties of composites, such as their tensile
properties, while the polymer matrix dominates the through-thickness or z-axis
properties, like compressive and flexural strength. Notably, any improvements to the
in-plane and in particular out-of-plane electrical conductivity require the addition of
a nanofiller. Therefore, the best way to improve the dominant properties of the
polymer matrix besides its thermal and electrical properties is to consolidate it with
conductive nanoparticles, such as GNPs. As such, the combination of carbon fibre
and a nanocomposite matrix to form three-phase composites or multiscale
composites can help to produce a new multiscale composites (MSCs) with better
flexural, thermal and electrical properties. This chapter will focus on the
characterisation of TPU70/CF and MSCs, incorporating 5 wt.% GNPM5, M15 and
M25 respectively. These characterisations entail the measurement of void content
and fibre volume fraction, dynamic mechanical thermal analysis (DMTA) of non-
annealed and annealed composites, and mechanical testing using tensile, flexural and
interlaminar shear stress (ILSS). Furthermore, the investigation of ILSS specimens
using SEM for different samples after testing was accomplished to identify the type
of failure that they exhibited. Finally, the measurements of in-plane and out-of-plane
electrical conductivity, thermal diffusivity using IR, and thermal conductivity were
also conducted.

181
5.2 Void Content and Fibre Volume Fraction

Both the fibre volume fraction and the void content have a great effect on the
performance of composites and multiscale composites [1]. Since the fibres control
the in-plane mechanical properties of the composite, the presence of voids is
considered one of the crucial manufacturing defects which reduce the desirable
properties of produced composites [1, 299, 300]. Interestingly, high content of voids
can cause weakened flexural modulus, flexural strength and interlaminar shear stress
of the composites [300, 301]. Voids can be formed during the lay-up of the
composite owing to the entrapment of mechanical air, or can be caused by moisture
absorbed by materials during their storing and processing [299, 302]. Voids can also
be formed during the manufacturing process of composite samples, especially when
the matrix possesses high viscosity itself, or as a result of incorporating nanofillers
[303]. The measurements of average fibre volume fraction (Vf), matrix volume
fraction (Vm) and void content (Vv) for all panels with different thicknesses that were
used during the present work are given in Table 5.1.

Table 5.1: Average properties of TPU70/CF and TPU70/GNPs/CF multiscale


composites
Carbon
Matrix Voids
Laminate Density of fibre
volume volume
Specimens/2mm thickness thickness composites volume
fractions fraction
(mm) (g/cm3), 𝜌𝑐 fractions
(Vm), (%) (Vv)%
(Vf), (%)
TPU70/CF 1.90  0.10 1.40  0.00 48.0  0.4 51.5  0.6 0.4  0.8
TPU70/5 wt.%GNPM5/CF 1.90  0.20 1.39  0.02 52.0  1.1 43.0  1.0 4.7  1.4
TPU70/5 wt.%GNPM15/CF 1.83  0.08 1.38  0.04 47.0  2.1 50.0  3.6 2.8  1.2
TPU70/5 wt.%GNPM25/CF 1.80  0.020 1.36  0.03 44.0  0.5 53.0  2.5 2.5  2.5
Specimens/3mm thickness
TPU70/CF 2.80  0.07 1.36  0.02 43.0  3.0 54.6  4.4 2.1  1.6
TPU70/5 wt.%GNPM5/CF 2.80  0.05 1.40  0.04 50.9  2.8 42.7  2.1 6.4  1.9
TPU70/5 wt.%GNPM15/CF 2.90  0.03 1.34  0.09 46.0  4.0 49.0  3.1 4.8  1.2
TPU70/5 wt.%GNPM25/CF 2.80  0.01 1.40  0.05 45.6  3.5 51.0  0.7 3.4  1.8
Specimens/4 mm thickness
TPU70/CF 3.70  0.10 1.28  0.10 46.0  2.5 50.0  2.4 2.1  1.7
TPU70/5 wt.%GNPM5/CF 3.80  0.10 1.27  0.05 47.0  5.0 45.0  5.0 7.6  1.6
TPU70/5 wt.%GNPM15/CF 3.70  0.02 1.32  0.01 44.0  0.5 50.0  0.6 6.0  0.5
TPU70/5 wt.%GNPM25/CF 3.80  0.03 1.33  0.06 44.0  1.5 52.0  4.0 4.0  0.2

182
From this table, the average carbon fibre volume fraction for all panels was measured
to lie between 43.0 and 52.0 %, and the void content was between 0.4 and 7.6 %. It
can be clearly observed that the void content increased with the presence of GNPs
due to the increase in viscosity of TPU-70 HS, as demonstrated previously (see
chapter four, section 4.3.5.2). This result suggests that the increase in viscosity
prevents the air from escaping so that it becomes trapped inside the matrix, causing
the formation of voids. It was demonstrated previously that the maximum increase in
viscosity attributed to the NCs matrices incorporated with GNPM15 and GNPM25
due to their higher surface area. In an unexpected way the maximum void content
and CF Vf was recorded to TPU70/CF composites incorporated with 5 wt.%
GNPM5, this might be because the GNPM5 possesses a smaller aspect ratio in
comparison with GNPM15 and M25, resulting in a decrease in TPU70-NC flow due
to agglomeration during melting. In addition, the decrease in matrix flow can reduce
the wettability between the matrix and the CF, resulting in an increase in fibre
volume fraction and void content, as reported previously [134].

The same findings were achieved by Wang et al. [303], albeit the MSCs is
completely different from the present one. The researchers investigated the effect of
two different types of GNPs (C750 and xGNPM5) on the morphology and
mechanical properties of GF/epoxy composite. They found that the void content
increased with the presence of nanoparticles and the maximum value was recorded
for the multiscale composites based on the C750 GNPs. The researchers attributed
their results to the small aspect ratio of C750, which was < 1m in diameter
compared with the GNPM5 diameter of > 1 m. As a result, the agglomeration of
C750 occurred and caused a reduction in the flow of the epoxy. These results give
substantial support to the results presented here.

5.3 DMTA Analysis

DMTA tests were performed in order to provide information on the stiffness of


composites, as well as on the transition and relaxation processes of the polymer
matrix along with the interfacial fibre-matrix and matrix-nanofiller adhesion. Figure
5.1 shows the results of log storage modulus (’) as a function of the temperature of
the TPU70/CF composite and MSCs before and after annealing. Table 5.2
summarises the data of ’ for the different samples at -100 °C, 25 °C, 100 °C and

183
also before and after annealing, which corresponds with Figure 5.1. The ’ of neat
TPU-70 HS was recorded previously at around 0.53 GPa at room temperature (see
Table A.3), indicating that the CF improved ’ by approximately 11 times. This
result may indicate the effective role of carbon fibre in carrying the stress from the
matrix, as well as demonstrating its higher modulus which causes this improvement
[304-307].

A) Non-Annealed B) Annealed
TPU70/CF TPU70/CF
TPU70/5wt%GNPM5/CF TPU70/5wt%GNPM5/CF
TPU70/5wt%GNPM15/CF TPU70/5wt%GNPM15/CF
Storage Modulus (GPa)

TPU70/5wt%GNPM25/CF TPU70/5wt%GNPM25/CF

10 10

1 1
-100 -50 0 50 100 -100 -50 0 50 100
Temperature (C) Temperature (C)

Figure 5.1: Storage modulus of A) non-annealed and B) annealed TPU70/CF


composites and MSCs incorporated with 5wt.% of GNPs based on their various sizes
of GNPs as a function of temperature
Table 5.2: Storage modulus value (E’) of TPU70/CF and TPU70-MSCs at -100, 25
and 100°C
Annealed/
Non-Annealed/Storage
Storage Modulus, E’,
Modulus, E’, (GPa)
Sample (GPa)
100
-100 °C 25 °C 100 °C -100 °C 25 °C
°C
10.0  6.0  2.2  8.3  5.2  2.0 
TPU70/CF
0.3 0.3 0.2 0.7 0.2 0.1
9.7  5.7  1.9  6.8  4.2  1.8 
TPU70/5wt%GNPM5/CF
0.5 0.4 0.1 0.3 0.3 0.2
14.0  9.0  2.6  11.3  6.9  2.5 
TPU70/5wt%GNPM15/CF
0.5 0.5 0.2 0.5 0.2 0.2
16.7  9.2  3.5 ± 13.5  8.0  3.0 
TPU70/5wt%GNPM25/CF
0.1 0.1 0.1 1.0 0.2 0.1
In general, all of the samples exhibit a linear trend in the glassy region before the
glass transition of the soft phase (TgSP) and decline sharply in the rubbery state after
TgSP due to the increase in chain mobility [251, 268]. Significantly, the incorporation
of 5 wt. % GNPM15 and 5 wt.% GNPM25 causes an enhancement in the ’ value of
TPU70/CF composites over the full temperature range before and after annealing.

184
Meanwhile, the addition of the same amount of GNPM5 causes a reduction in ’ of
TPU70/CF at all temperatures. For example, at room temperature (see Table 5.2) the
’ values of non-annealed samples added to 5 wt.% GNPM15 and GNPM25 increase
to 9 GPa and 9.2 GPa respectively from 6 GPa of TPU70/CF, whereas the ’ values
of MSC samples with GNPM5 decrease to become 5.7 GPa compared to TPU70/CF.
A similar improvement is observed in both glassy states, such as at -100 °C, and in a
rubbery state like 100 °C, albeit that the glassy state (-100 °C) shows the greater
enhancement by 66 % and 40 % for GNPM25 and GNPM15 respectively in
comparison with TPU70/CF laminates.

The disparity between these results may be due to a number of factors. Firstly, the
good dispersion of GNPM15 and GNPM25 inside the matrix is likely to be better
than for GNPM5, although this suggestion could not be verified using SEM or
optical microscopic analysis, as demonstrated previously (see sections 4.3.3.1 and
4.3.3.2). Both SEM and OM images showed approximately the same characteristics
for all GNP-NCs samples. Secondly, the GNPM15 and GNPM25, in particular,
impart good interfacial interaction with cyanate groups in TPU-70 HS structure as
demonstrated previously in chapter four (see sections 4.3.1 and 4.3.2) on one side
and NCs-CF on the other side due to their higher aspect ratio. The good interfacial
adhesion would impart better stress transfer between the fibre and nanofiller
reinforcement, as demonstrated previously [9, 303]. In contrast, the MSCs
incorporating GNPM5 would reduce the interfacial interaction with the matrix due to
its smaller size, thereby reducing the quality of stress transfer between the matrix and
the CF.

In addition, this type of sample contains a higher percentage of voids in comparison


with other samples, as summarised in Table 5.1, and the void content has a great
effect on composite stiffness. As such, the reduction in ’ of TPU70/CF with
GNPM5 can be attributed to this characteristic. Furthermore, the agglomeration of
GNPM5 during melting due to its small size has a detrimental influence on fibre
impregnation [308]. The lack of fibre impregnation would reduce the matrix-fibre
interfacial adhesion and then increase the internal porosity, which can offer an
explanation for the high void content of the TPU70/5 wt. %GNPM5/CF composite.
These results are consistent with DMTA results obtained from studying the effect of
adding two different sizes of GNPs to epoxy/GF [303]. It was demonstrated that the

185
addition of GNPM5 gives a better improvement in ’ and Tg of epoxy/GF than GNP
C750 due to its higher surface area and good interfacial interaction. Overall, the
present results are compatible with the DMTA results of PP/nano-clay/CF [304],
which revealed that the addition of small amounts of nano-clay to PP/CF
considerably improved its storage modulus, as supported by other DMTA results
[307].

After annealing, as shown in Table 5.2 and Figure 5.1 B, the values of log E’ of all
four samples decrease from their value before annealing, despite the samples
displaying the same trend across the whole temperature range. For instance, in the
glassy region at -100 °C exactly, the value of E’ of TPU70/CF, TPU70/5
wt.%GNPM5/CF, TPU70/5 wt.%GNPM15/CF and TPU70/5 wt.%GNPM25/CF
composites decrease to 8.3 GPa, 6.8 GPa, 11.3 GPa and 13.5 GPa respectively, when
compared to their value before annealing. Similar trends in E’ value reduction occurs
in the rubbery state at 25 °C and 100 °C. This reduction is likely to be caused by all
GNP types increasing the demixing of SS and HS, due to their hindrance of the
mobility of the polymer chain during annealing. This process results in a reduction in
the microcrystalline hard domain, followed by a reduction in microphase separation,
as demonstrated in Chapter four (see section 4.3.4.3) and in other studies [82, 253].
Alternatively, it may be attributed to the long duration of the annealing treatment,
which not only reduces the interfacial interaction between GNPs and the TPU matrix
but also affects the adhesion between CF and TPU-70 HS or NCs, causing the
stiffness of the composites to decrease and subsequently E’ to decrease as well.
However, these findings are incompatible with previous DMTA results for TPU-70
NCs, which were explained in detail in section 4.3.4.3 and are consistent with the
results reported by Albozahid [46].

DMTA is considered to be the best method for determining the glass transition
temperature of polymers and to study its influence by combining polymer matrices
with nanofillers and fibre. Figure 5.2 A and B illustrates tan  curves of TPU70/CF
and TPU70/GNPs/CF composites before and after annealing. Table 5.3 summarises
the corresponding average Tg of the soft phase and hard phase of TPU70/CF and
TPU70/GNPs/CF composites before and after annealing as well.

186
A) Non-Annealed B) Annealed
TPU70/CF 0.14
TgHP TPU70/CF
TPU70/5wt%GNPM5/CF TPU70/5wt%GNPM5/CF TgHP
0.15 TPU70/5wt%GNPM15/CF
0.12 TPU70/5wt%GNPM15/CF
TPU70/5wt%GNPM25/CF
TPU70/5wt%GNPM25/CF

0.10

Tan 
Tan 

0.10 Second TgSP TgSP


Relaxation 0.08
Second
Relaxation
0.06
0.05

0.04

0.00 0.02
-100 -75 -50 -25 0 25 50 75 100 125 -100 -75 -50 -25 0 25 50 75 100 125
Temperature (C) Temperature (C)

Figure 5.2: Tan delta (loss factor) of TPU70/CF laminate and multiscale composites
A) before annealing and B) after annealing
Table 5.3: Average glass transition temperature of soft and hard phases before and
after annealing of TPU70/CF laminate and multiscale composites

Non-annealed Annealed Non-annealed Annealed


Sample TgSP TgSP TgHP TgHP
TPU70/CF -15.7  1.8 -14.0  0.4 80.1  1.6 78.5  0.9
TPU70/5 wt.%GNPM5/CF - -14.8 1.2 90.6  0.6 78.3  0.6
TPU70/5 wt.%GNPM15/CF - -15.6  1.4 91.0  0.7 78.2  0.5
TPU70/5 wt.%GNPM25/CF -13.5  0.5 -15.1  0.1 92.0  0.4 79.3  1.4
Overall, the tan  curves of TPU70/CF and MSCs exhibit three peaks similar to those
of neat TPU-70 HS and TPU70-NCs (see 4.3.4.3). The smaller and lower peak
represents the second relaxation, located at -73 °C, which is approximately the same
value for all samples, regardless of whether they are non-annealed or annealed. The
second peak represents the glass transition temperature of the soft phase (TgSP) which
occurs at a slightly lower temperature for the non-annealed samples as a result of the
inclusion of GNPM25, while it disappears for the MSCs incorporating GNPM5 and
M15. The TgSP maintains almost the same value after annealing for TPU70/CF and
TPU70/GNPM25/CF, and it can be detected for GNPM5 and GNPM15. In contrast,
the higher peak represents the glass transition temperature of the hard phase (TgHP),
which shifts to higher temperatures by consolidating with GNPs for the non-annealed
samples and to the lower temperature of the annealed samples.

The interfacial interaction between the reinforcement and a polymer has a great
effect on transition temperature on both the macro and micro scale [309]. The
inclusion of fillers in a polymer structure can reduce the free volume and inhibit
polymer chain mobility, reflected in an increase in Tg. Thus, the inclusion of CF in
TPU-70 HS causes the TgHP to increase from 67 °C for neat TPU (see section 4.3.4.3)

187
to 80 °C; this improvement increases with the presence of GNPs. These results are
compatible with those reported by Diez-Pascual et al. [9], who found that the
addition of CNT to PEEK/GF showed an enhancement in Tg compared with the
PEEK/GF system; their results are consistent with another previous study [309].

The tan  or damping factor in the transition regions determines the amount of
energy exploited to deform materials which is directly dissipated into heat [9, 114].
Therefore, it is found that composites with better interfacial adhesion between the
reinforcement and the resin have a lower tan  peak value than those with poor
adhesion [310, 311]. This consequence is ascribed to the low amount of heat that is
dissipated due to the internal friction between the resin and the reinforcement. In
spite of the similarity in the shifting TgHP of samples consolidated with GNPs, their
tan  peaks values are different. This inequality in tan  peak values is due to the
difference in aspect ratio and surface area of GNP types. For instance, for MSCs
incorporating GNPM25, their highest value of tan  peak is lower than that of the
TPU70/CF samples and other MSC samples which are incorporated with GNPM15
and M5. This result can be connected to the ability of GNPM25 to form more
interfacial interactions with the matrix, as confirmed in the previous chapter, which
gives more hindrance to the molecular movement of polymer chains. In addition, the
decrease in value of tan  maximum peaks could be attributed to the decrease in a
mechanical loss that overcomes the internal friction between molecular chains due to
the addition of GNPM25. Furthermore, the void content percentage in
TPU70/GNPM5/CF and TPU70/GNPM15/CF samples could affect the fibre
impregnation process, as mentioned earlier in this section, leading to another factor
that may produce the effect of restricting chain mobility. These results are also
consistent with a previous study of epoxy/GNP/GF [303], although the materials are
different from the present one.

After annealing, the TgSP and TgHP data in Figure 5.2 and Table 5.3 are almost the
same for the TPU70/CF and TPU70/GNPs/CF samples, regardless of their types.
However, there is a marginal change in the TgSP value of the annealed samples,
specifically of its value of TPU70/CF before annealing and of TPU70/GNPM25/CF
composite; these results are unknown for MSCs incorporating GNPM5 and M15, as
they were undetectable before annealing. The TgHP of all samples exhibits a
significant reduction in value before annealing. This result could indicate that the

188
annealing treatment helps to restack the GNPs with each other by destroying the
weak bonds between them and HS in the TPU-70 HS matrix causes an obstacle to the
formation of a microcrystalline hard domain. This phenomenon can lead to a
reduction in the degree of microphase separation, as explained earlier in this section.

5.4 Mechanical Testing

5.4.1 Tensile Test

The typical tensile stress-strain curves of TPU70/CF and TPU70/5 wt.%GNPM5,


TPU70/5 wt.%GNPM15 and TPU70/5 wt.%GNPM25/CF-MSCs before and after
annealing are illustrated in Figure 5.3. The corresponding ultimate tensile strength
(𝜎𝑢𝑙𝑡 ), modulus of elasticity (𝐸), and the percentage of elongation at break or final
strain% at failure (𝜀𝑓 %) are plotted in Figure 5.4. In general, all of the samples
exhibit a brittle fracture in both cases, before and after annealing, without a clear
point of yielding. After the addition of GNPs, an obvious change in the behaviour of
the stress-strain curve of TPU70/CF can be recognised.

A) Non-Annealed B) Annealed
600 600
TPU70/CF
TPU70/5wt.%GNPM5/CF
500
Tensile Stress (MPa)

TPU70/5wt.%GNPM15/CF 500
TPU70/5wt.%GNPM25/CF
400 400

300 300

200 200
TPU70/CF
100 TPU70/5wt.%GNPM5/CF
100
TPU70/5wt.%GNPM15/CF
TPU70/5wt.%GNPM25/CF
0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Axial Strain % (mm/mm) Axial Strain % (mm/mm)

Figure 5.3: Typical tensile stress-strain curves of TPU70/CF and TPU70/5 wt.
%GNPM5, TPU70/5 wt.%GNPM15 and TPU70/5 wt.%GNPM25/CF-MSCs A)
before annealing and B) after annealing
It can be observed in Figure 5.3 A that the reference samples of TPU70/CF are
exposed to a rapid linear increase in stress with strain until reaching the fracture
point (𝜎𝑢𝑙𝑡 ), which is (36160) MPa. While, with the presence of 5 wt.% GNPM5,
the trend of stress curve increases gradually with strain until breaking at a point
lower than that of the TPU70/CF samples, around (28013) MPa. On the contrary,
the addition of 5 wt.% GNPM15 and 5 wt.% GNPM25 makes the trend of stress-

189
strain curves slightly more rapid, breaking at a point higher than TPU70/CF,
achieving a value of 𝜎𝑢𝑙𝑡 at around 46927 MPa and 55170 MPa respectively.

Before annealing, the gradient of the straight line in Figure 5.3 A and B, or the
elastic modulus (𝐸), shows a marginal increase with the addition of 5 wt. %
GNPM15 and M25 and a slight decrease with the presence of GNPM5 compared to
the reference sample (TPU70/CF), as illustrated in Figure 5.4 A. However, after
annealing, the tensile stress-strain curve of TPU70/CF becomes slightly higher and
sharper than that of the MSCs, which indicates an increase in its elastic modulus and
its stiffness. This gentle improvement in E of TPU70/CF and tensile strength and
percentage of elongation at break after annealing might refer to the promotion of the
formation of microcrystalline hard domain and increased phase separation during
annealing [46, 47, 82]. The improvement in phase separation leads to an
enhancement in the mechanical properties of neat TPU. As a result, the E, 𝜎𝑢𝑙𝑡 and
elongation at break of TPU70/CF are enhancement after annealing. For instance, the
value of E of TPU70/CF increases from 41.0  7.7 GPa before annealing to 46.0 
5.6 GPa after annealing.

Normally, in a tensile test, the fibre has a more dominant impact on the in-plane
mechanical properties than the matrix. Nevertheless, this tendency does not eliminate
the effect of the matrix, which plays a crucial role by transferring load between
reinforcement fibres. Good adhesion between fibres and a matrix can increase the
ability of the matrix to transfer load between composite layers, thereby increasing the
tensile properties. For NC matrices, it has been believed that the addition of GNPs,
regardless of their sizes, would increase the mechanical properties of a matrix, as
reported in previous studies [27, 312-315]. In fact, there are other factors which can
affect this outcome: the interfacial adhesion between fibres and NC matrices, and the
dispersion of nanofiller within a polymer matrix [111]. Therefore, adding 5 wt.%
GNP M5 causes a reduction in 𝜎𝑢𝑙𝑡 and a slight decrease in E, which indicate the
failure of TPU70/GNPM5 to transfer the load, leading microcracks to form in the
matrix. In addition, the stress concentration at the interface between GNPM5 and the
TPU-70 HS matrix may increase owing to the poor interaction between them,
resulting in a reduction in GNPM5/NCs to transfer the load and then fail in a low
𝜎𝑢𝑙𝑡 and elongation at break compared to TPU70/CF, as shown in Figure 5.4 A and
B.

190
(A)
600 Non-Annealed Annealed 60
Non-Annealed Annealed
525
50
Tensile Strength (MPa)

Elastic Modulus (GPa)


450
40
375

300 30

225
20
150
10
75

0 0
F F F
PM
5/C 15/C 25/C
/CF GN NPM NPM
70 t.% G G
TPU t.% t.%
U70/5w 70/5w 70/5w
TP TPU TPU

(B)
1.8
Non-Annealed
1.6 Annealed
Elongation at break % (mm/mm)

1.4

1.2

1.0

0.8

0.6

0.4

0.2

0.0
F F F
P M5/C M15/C M25/C
F GN P P
70/C wt.%
GN GN
TPU 0/5 0/5wt.% 0/5wt.%
7 7 7
TPU TPU TPU

Figure 5.4: A) Ultimate tensile strength and elastic modulus and B) elongation at
break % of TPU70/CF and TPU70/GNPs/CF-MSCs before and after annealing
In contrast, the MSCs incorporating GNPM15 and GNPM25 show a slight
improvement in elastic modulus and the elongation at the break before annealing.
This outcome can be ascribed to the good interfacial interaction between GNPM15
and GNPM25 with the HS in the TPU structure, as demonstrated previously in
chapter four (see sections 4.3.1 and 4.3.2). Hence, the good interaction between the

191
GNPs and the TPU matrix may transfer the load and carry some of it along with the
CF, instead of enhancing the stress concentration at the interface between GNPs and
TPU-70 HS matrix like MSCs incorporated with GNPM5.

On the other hand, the reduction in the elongation at break with the addition of
GNPM5 brings about the reduction in the plastic deformation of the TPU-70HS
matrix. This result is due to the restriction which occurs with the agglomeration of
GNPM5 resulting from weak bonding with the matrix, as reported in previous studies
[316, 317]. Nonetheless, the MSCs which included GNPM15 and GNPM25
exhibited a slight increase in the elongation at break in comparison with TPU70/CF
by 12% and 36% respectively, as displayed in Figure 5.4 B. This outcome is ascribed
to the reason suggested previously in relation to the good interfacial interaction of
GNPM15 and GNM25, which reduces the stress concentration and enhances the
ductility of the MSCs.

The current findings are in agreement with the behaviour of PEEK/SWCNT/glass


fibre composites reported by Diez-Pascual et al. [9] and mentioned previously in
literature review in section 2.4.4.1. The researchers found that the addition of 1 wt.%
of non-wrapped and wrapped SWCNTs in polysulfone to PEEK/GF causes a slight
increase in E (around 14% and 16% respectively). Conversely, the elongation at
break decreases for non-wrapped SWCNT/MSCs and moderately increases for
wrapped SWCNTs due to the weak bonding between non-wrapped SWCNTs, which
increases the resistance to polymer chain, leading to a decrease in the extent of
plastic deformation. On the contrary, the wrapped SWCNTs form a strong interfacial
interaction, causing an enhancement in the ductility of the PEEK composites, as
confirmed in a previous study [317]. The current results are compatible with other
previous observation in the literature review which reported by Gojny et al. [121].
The researchers revealed that the incorporation of CNTs with epoxy/GF result in a
marginal improvement in its in-plane mechanical properties.

As mentioned previously, all of the tensile properties of TPU70/CF show


improvement after annealing. Nevertheless, the tensile strength decreases for all
GNP/MSC types, while the elongation at break increases slightly for MSCs,
including GNPM5 and GNPM15, and decreases marginally for MSCs incorporating
GNPM25. In fact, the aspects of MSC behaviour observed after annealing are very
complicated to explain; it is likely that the reduction in hydrogen bonding (HB)

192
between GNPM25 and the HS of the TPU molecular chain granted the nanoparticles
an opportunity to restack, resulting in agglomeration. The occurrence of
agglomeration of the composites during the tensile test obstructed the drawing of
polymer chains, resulting in a reduction in the extent of plastic deformation besides a
reduction in the stiffness of the composites. On the other hand, the GNPM5 and
GNPM15 might work as a nucleate agent which helps to increase the
microcrystalline HD. In turn, this might increase the microphase separation of the
TPU matrix, which is believed to improve the mechanical properties of TPU70/NC
matrices, as demonstrated previously in Chapter 4 and in other studies [46, 253, 254].

In terms of fracture type, Figure 5.5 shows the visual observation of fracture damage
that found in the TPU70/CF laminates and MSCs after tensile test.

(A)

Localize damage near tabs

(B)

TPU70/GNPM5/CF interfacial failure

(C)

Localize damage

(D)

Figure 5.5: Failure pattern of A) TPU70/CF laminates, B) TPU70/GNPM5/CF, C)


TPU70/GNPM15/CF and D) TPU70/GNPM25/CF multiscale composites after
tensile test

193
It can be observed from Figure 5.5 A that the matrix crack near the tabs without
longitudinal splitting or fibre pull-out is found in TPU70/CF laminate. On the other
hand, the MSCs combined with GNPM5 exhibits a mixed fracture comprised of a
longitudinal splitting, interfacial failure between fibre and matrix and fibre pull-out
as shown in Figure 5.5 B. This mixed failure indicates a weak interfacial adhesion
between GNPM5-NCs and CF as interpretation earlier in this section. Both MSCs
combined with GNPM15 and GNPM25 exhibit a failure similar to TPU70/CF
laminate in spite of fewer observation of fibre pull-out. The absence of fibre-matrix
interfacial failure confirms the previous hypothesis of good adhesion between
GNPM15 and GNPM25-NCs and CF.

5.4.1.1 Theoretical Micromechanical Analysis Model


The elastic modulus can be predicted for polymer/nanofiller/fibre MSCs using a
modified rule of the mixture model (RoM) as a simple model for predicting the
elastic modulus of a two-phase system. The RoM is described in equation (5.1):

𝑬𝒄 = 𝒆 𝑬𝒇 𝑽𝒇 + 𝑬𝒎 𝑽𝒎 ......... ( 5.1)

where EC, Ef and Em are the elastic moduli for composites, fibre and matrix
respectively, and e is the fibre efficiency factor, which equals 0.5 for a continuous
bidirectional fibre [318]. It should be taken into consideration that this model was
built on the assumptions that the bonding between the fibre and the matrix are perfect
and that the composites are free from voids, as well as the understanding that the
properties of both the matrix and the fibre are isotropic [319]. E for the CF was taken
as 250 GPa from Table 3.7, as reported by Sigmatex company [151].

Em can be considered as the average value of the elastic modulus of TPU-70 HS,
which was measured experimentally in Chapter 4 and was found to be approximately
equal to 0.2 GPa, while for NCs matrices can be estimated using Krenchel’s rule of
mixture equation for discontinuous reinforcement.

𝑬𝑵𝑪𝒔 = 𝒍 𝟎 𝑽𝑷 𝑬𝑷 + 𝑽𝒎 𝑬𝒎 ......... ( 5.2)

where 𝑉𝑃 and 𝑉𝑚 are the volume fraction of the fillers and the matrix respectively; 𝐸𝑃
is the elastic modulus of the filler; Em is the elastic modulus of the matrix, which is
0.2 GPa of TPU-70 HS; 0 is the orientation factor, which is 1 for aligned
nanoplatelets and 8/15 for nanoplatelets in a 3D random orientation, which was

194
suggested and proven recently by Li et al. [320, 321] and adopted in this study; and
𝑙 is the length efficiency factor that can be calculated using the following equation
[322, 323]:
𝒍
𝒍 = 𝟏 − (𝟐𝒍𝒄 ) ......... ( 5.3)

where 𝑙𝑐 is the critical length (3m was adopted as the required graphene flake
length so as to obtain better reinforcement in the polymer NCs [324]), and l is the
filler length, which was 5, 15, and 25 m for GNPM5, GNPM15 and GNPM25,
respectively.

For the E of GNPs, it has been demonstrated that the in-plane tensile modulus of a
single graphene sheet is 1060 GPa [325]. However, it is well known that the GNPs
consist of several graphene sheets stacked together. When the NCs were subjected to
tensile stress, this stress transferred from the matrix to the GNP particles, and the
weak Van der Waals bonds between the graphene sheets in the GNP structure were
more likely to fail than the graphitic carbon-carbon bonds within the graphene sheet,
leading to further exfoliation in the GNPs. Therefore, the E of the GNPs was equal to
the elastic modulus of exfoliated graphite in the z-axis (through-thickness), which
measured 36.5 GPa [326, 327].

Figure 5.6 illustrates the comparison between predicted elastic modulus using the
RoM model and the experimental measurements of the TPU70/CF and MSCs. It can
be clearly observed that the presence of GNPs in TPU70/CF composites causes a
slight decrease in the predicted E of the MSCs combined with GNPM15 and
GNPM25, and a gentle increase in that of the MSCs mixed with GNPM5, the
opposite of the results of the experimental work. This behaviour of the theoretical
model may be attributed to the discrepancy in Vf for the different MSC samples; the
MSCs consolidated with GNPM5 yielded the higher value (52%), while the
GNPM25-MSCs gave the lower value (44%). However, the value of E in the
predicting model is higher than that of the experimental data measurements. This
result may have arisen from a number of factors. For example, the waviness of CF is
likely to decrease its effective elastic modulus and reduce its capability to resist the
applied load, as well as enhancing composite stiffness. Furthermore, the dispersion
of nanofillers and the interfacial interaction between them and the TPU-70 HS matrix
are crucial factors that influence the interfacial stress transfer between fibre

195
reinforcements. The differences between the two results can be seen in the MSCs;
this is because the RoM eliminates the effect of interfacial interaction between the
NCs and CF.

80 Rule of Mixture
Experimental
70
65
Elastic Modulus (GPa)

60 59
60
55
50

40

30

20

10

0
CF F F
P M5/ M 15/C M25/C
F N P P
70/C wt.%
G GN GN
TPU 0/5 0/ 5wt.% 0/5wt.%
7 7 7
TPU TPU TPU

Figure 5.6: Comparison between predicted elastic modulus and results of


experimental measurement of TPU70/CF and MSCs as functions of different
matrices
The same agreement was reported by Pascual et al. [9] using the RoM model. They
suggested that the discrepancies between the theoretical and experimental values of E
of the MSCs could be ascribed to the effect of the dispersion of the SWCNTs and to
the interaction between the PEEK matrix and the two filler phases, SWCNTs and
GF, which in turn influence the transfer stress between them.

According to the results of the theoretical model concerning the effect of Vf on the
tensile properties, it becomes essential to normalise the tensile modulus and tensile
strength to the higher Vf, which is 52% for samples with 2 mm thickness (see Table
5.1). The similar manner of normalisation to the mechanical testing results had been
carried out by Junid [10] and Liu [1] of MSCs based on the epoxy/CF and
incorporated with GNPs and CNTs respectively. Therefore, Table 5.4 and Table 5.5
show a comparison between the actual and normalised values of  and ult of
TPU70/CF and MSCs before and after annealing, respectively. It is noteworthy to

196
observe that taking the effect of Vf by normalising the tensile data with the higher
value of Vf results in a value that is roughly double for all samples.

Table 5.4: Comparison between actual and normalised values of tensile properties
of TPU70/CF and MSCs according to the high Vf percentage (Vf=52%) before
annealing
Sample Actual value/non-annealed Normalised value/non-annealed
E(GPa) ult (MPa) E(GPa) ult (MPa)
TPU70/CF 41.0 ± 3.0 361.0 ± 60.0 79.2 ± 6.0 694.0 ± 141.0
TPU70/5wt.% GNPM5 35.3 ± 1.4 280.3 ± 13.3 67.5 ± 3.4 539.0 ± 31.3
TPU70/5wt.% GNPM15 46.2 ± 2.3 469.0 ± 27.3 88.8 ± 5.2 901.9 ± 60.7
TPU70/5wt.% GNPM25 48.8 ± 3.1 551.3 ± 70.3 93.9 ± 7.3 1060.3 ± 166.0

Table 5.5: Comparison between actual and normalised values of tensile properties of
TPU70/CF and MSCs according to the high Vf percentage (Vf=52%) after annealing
Sample Actual value/annealed Normalised value/annealed
E(GPa) ult (MPa) E(GPa) ult (MPa)
TPU70/CF 46.4 ± 5.6 440.7 ± 39.3 89.3 ± 13.2 847.4 ± 92.7
TPU70/5wt.% GNPM5 34.6 ± 2.1 303.8 ± 41.7 66.6 ± 5.6 584.1 ± 92.6
TPU70/5wt.% GNPM15 45.6 ± 2.0 482.3 ± 43.2 87.8 ± 4.6 927.4 ± 95.9
TPU70/5wt.%GNPM25 43.3 ± 3.2 559.0 ± 62.5 83.2 ± 7.8 1075.0 ± 146.0

5.4.2 Flexural Test

Typical stress-strain curves of TPU70/CF and MSCs from a flexural test before and
after annealing are shown in Figure 5.7. The TPU70/CF exhibits a ductile failure
mode when the curve starts to increase linearly until yielding at peak stress, after
which it drops gradually. The same curve trend of TPU70/CF is observed after
annealing, but with a smooth failure without a sharp drop after a maximum load
point. The presence of 5 wt.%GNPM5, 5 wt.%GNPM15 and 5 wt.%GNPM25 in the
TPU70/CF causes an obvious increase in the slope of the linear deformation zone
and the maximum flexural stress of MSCs curves, suggesting that the presence of
GNPs increases the stiffness of the composites. The curves of the MSCs follow the
same trends of the TPU70/CF samples, decreasing gradually or fluctuating (such as
GNPM15-MSCs after annealing) after reaching a maximum peak stress. This type of
progressive failure is comprised of fibre failure, matrix-fibre debonding, and
delamination. Nevertheless, non-annealed MSC samples with GNPM25 show an
irregularity after linear deformation up to 0.003 mm/mm, which might be attributed
to fibre breakage. The curve then continues to increase until a peak point, then

197
decreases gradually without a sudden break. It is believed that the slight reduction in
the load after reaching a maximum peak stress for all of the TPU70/CF and MSCs
samples is due to the fibre breakage accompanied by a gentle increase in strength,
which is ascribed to matrix cracking and local delamination [328].

A) Non-annealed B) Annealed
160 160

140 140
Flexural Stress (MPa)

120 120

100 100

80 80

60 60
TPU70/CF
TPU70/CF TPU70/5wt.%GNPM5/CF
40 40
TPU70/5wt.%GNPM5/CF TPU70/5wt.%GNPM15/CF
TPU70/5wt.%GNPM15/CF TPU70/5wt.%GNPM25/CF
20 20
TPU70/5wt.%GNPM25/CF
0 0
0.000 0.003 0.006 0.009 0.012 0.015 0.018 0.000 0.004 0.008 0.012 0.016 0.020 0.024
Flexural Strain (mm/mm) Flexural Strain (mm/mm)

Figure 5.7: Typical flexural stress-strain curves of TPU70/CF and MSCs A) before
and B) after annealing
As shown in Table 5.1, the dissimilarities in the Vf and void content of TPU70/CF
and MSCs have a great effect on the mechanical properties, hence the flexural
strength and flexural modulus are normalised. As with the tensile properties, the
normalisation was conducted in accordance with the higher value of Vf (52%) for the
TPU70/5 wt.% GNPM5/CF samples with 2 mm thickness.

The normalised flexural strength (FS) and flexural modulus (FM) of the TPU70/CF
and MSCs where Vf = 0.52 are presented in Figure 5.8. The corresponding value is
tabulated in Table B.1and Table B.2 in Appendix B. It can be clearly seen that the
flexural strength and flexural modulus of the TPU70/CF composites are improved by
the addition of GNPs, regardless of their size, as a result of the reinforcing effect of
the high modulus of the GNPs. Specifically, after adding 5 wt.% GNPM5, 5 wt.%
GNPM15 and 5 wt.% GNPM25 and before annealing, the flexural modulus increases
by 32%, 48% and 105%, from 27  2.5 GPa to 372.8 GPa, 41  7.8 GPa and 57 
3.8 GPa, respectively. Similarly, to the FM, the FS of the TPU70/CF composites
rises from 235  17.4 MPa to 253  8 MPa, 262  12 MPa and 295  13 MPa for
MSCs incorporating GNPM5, GNPM15 and GNPM25 respectively. The increases
are 8%, 11% and 26% higher than those of the FS of the TPU70/CF composites.
198
400 80
Non-Annealed Annealed

Normalised Flexural Modulus (GPa)


Normalised Flexural Strength (MPa)
360 Non-Annealed Annealed 70
320
60
280
50
240

200 40

160
30
120
20
80
10
40

0 0
5/C
F /CF /CF
C F NPM NP M15 NP M25
70/ G G G
TPU t.% t.% t.%
U70/5w 70/5w 70/5w
TP TP U TP U

Figure 5.8: Normalised flexural strength and flexural modulus for TPU70/CF and
MSCs before and after annealing the normalisation have been done according to the
high Vf percentage (Vf=52%)
It can also be observed from the figure that the FS and FM values of TPU70/5 wt.%
GNPM25/CF are higher than those of GNPM5 and GNPM15. The robust ability of
GNPM25 to cause this improvement is mainly ascribed to its larger size and good
interfacial interaction with the TPU-70 HS and/or NC matrix with CF, as proven
previously in the tensile test. Similar behaviour was reported by Wang et al. [303],
who used two different GNP sizes, GNPM5 and GNP-C750, incorporating them with
an epoxy matrix and GF composites. They demonstrated that the FM of epoxy/GF
composites mixed with GNPM5 increased by 26.3% compared with 11.5% of
epoxy/GF composites with GNP-C750. They ascribed these differences in
improvement to the larger aspect ratio and good dispersion of GNPM5, which might
produce MSCs with strong interfacial interactions between GNPM5 and the epoxy
matrix, on the one hand, and NCs matrices and GF on the other hand.

It is noticeable that the addition of GNPs to TPU70/CF composites results in a


remarkable improvement in FM and FS, while adding GNPs to TPU70/CF
composites in the tensile test caused only a slight increase in tensile properties, as
mentioned previously. This finding can be attributed to the reinforcement effect of
GNPs in the out-of-plane direction when the matrix has the dominant mechanical
properties. From the literature review, Kim et al. [130] reported that the modification

199
of epoxy/CF composites by the dispersion of 0.3 wt.% CNT in epoxy resin had a
slight effect on tensile properties, whereas the FM and FS increased by 11.6% and
18% respectively in comparison with epoxy/CF composites. Similar findings were
reported by Diez-Pascual et al. [9]. These earlier studies give good agreement with
the results presented here. Due to the limiting work on TPU/GNPs/CF composites
system or limiting works on MSCs based on the TPs matrix and reinforced by CF,
different MSCs are used for the purpose of comparison. Table 5.6 shows the current
normalised flexural results of MSCs incorporating GNPM25 with the flexural test
results of other composites combined with various nanofillers.

Table 5.6: Comparison between the FM and FS of current results of TPU70/CF and
MSCs before annealing with other composites incorporating various nanofillers
Composites Matrix Fibre Fillers Content FM FS Ref.
% (GPa) (MPa)
CFRP/GNP epoxy carbon GNPs 1 vol 63.8  679  [111]
1.4 7
CFRP/GNPs epoxy carbon GNPs 0.5 wt. 77  4.0 1260 [131]
 62
GFRP/GNP epoxy glass GNP- 5 wt. 26.0  360  [303]
5m 2.0 40
GFRP/GNP-NH2 epoxy glass GNP- 12 wt. 32.0  520  [329]
NH2 1.5 50
GFRP/laser- PEEK glass Laser- 1.0 wt. 22.8  368  [9]
SWCNT+PEES SWCNT 0.3 1
CFRP/GNPM25 TPU- carbon GNPM2 5 wt. 57.3  295  Present
70 HS 5 3.8 13 work
Regardless of the manufacturing process adopted to produce the composites listed
above and of the conditions and method followed during the flexural test, the present
results yield a moderate value of FM and a lower value of FS. This outcome might be
attributed to the fact that thermoplastic matrices (TPU-70 HS), in general, have lower
flexural properties than thermoset matrices (epoxy matrices).

Furthermore, from Figure 5.8 it can be observed that the FM and FS of TPU70/CF
and MSCs incorporated with GNPM5 and GNPM15 improved after annealing, while
the MSCs with GNPM25 exhibited a slight reduction. This outcome is consistent
with the tensile test results of the TPU70/CF composites and MSCs after annealing,
with the exception that the GNPM25-MSCs has a reduction in FS, while its tensile
strength exhibited a slight increase. Therefore, the same interpretation of this
reduction suggested previously can be applied in this situation. The improvement in
the FS and FM of MSCs combined with GNPM5 and GNPM15 is particularly

200
noteworthy, displaying increases from 37  2.8 GPa and 41.4  8 GPa before
annealing for GNPM5 and GNPM15-MSCs to 46  7.8 GPa and 51.8  4 GPa,
respectively. It is believed that this enhancement can be attributed to GNPM5 and
GNPM15 acting as nucleating agents during annealing, leading to the accelerated
formation of microcrystalline HD and increasing the phase separation between SD
and HD, as has been demonstrated and explained in the present study and other
previous studies [46, 253, 254].

5.4.3 Interlaminar Shear Strength (ILSS)

Figure 5.9 A and B show the curves of stress-strain results after a short-beam shear
test of the TPU70/CF composites and TPU70/5 wt.%GNPs/CF/MSCs, before and
after annealing. The trend of the TPU70/GNPs/CF/MSC curves before and after
annealing show a clear peak where they begin to increase linearly with the strain
until they reach a maximum point, then drop to a certain level and gradually shift
from the peak. These results might suggest that the maximum point represents the
interlaminar shear failure due to shear deformation, which causes a de-bonding or
delamination in the composite structure, as proven in the SEM images (see Figure
5.11 and Figure 5.12).

A) Non-Annealed B) Annealed
28 28

24 24

20 20
Stress (MPa)

16 16

12 12
TPU70/CF TPU70/CF
8 TPU70/5wt%GNPM5/CF 8 TPU70/5wt%GNPM5/CF
TPU70/5wt%GNPM15/CF TPU70/5wt%GNPM15/CF
4 TPU70/5wt%GNPM25/CF 4 TPU70/5wt%GNPM25/CF

0 0
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Strain % (mm/mm) Strain % (mm/mm)

Figure 5.9: Stress-strain curves after short-beam shear test of TPU70/CF laminates
and TPU70/5 wt.%GNPs/CF multiscale composite A) before annealing and B) after
annealing
Moreover, this behaviour of the curves obtained for the MSCs likely indicates the
heterogeneity of the NCs due to a degree of aggregation of the GNPs. On the
contrary, the TPU70/CF samples do not show any failure due to shear deformation;
their stress-strain curves rise linearly until reaching the highest point, remaining

201
steady for a while before displaying a final upward trend. The behaviour of the
curves of the TPU70/CF samples implies that these samples neither failed in shear
nor in debonding, but that they failed due to compression and fibre breakage, as
shown in the SEM images in Figure 5.11 A.

It is notable that for the MSC samples the ultimate ILSS value is lower than that of
the TPU70/CF samples and the lowest value is ascribed to the MSCs combined with
GNPM5. It is believed that the ILSS value was significantly influenced by the
percentage of voids in the composite samples during the manufacturing and
preparation process [300, 330-332]. Costa et al. [330] revealed that epoxy/CF
composites with a percentage of voids greater than 0.9% caused a marked decline in
their ILSS values. Their findings were compatible with those of Hong-Yan et al.
[333] and Liu [302], who revealed that an increase in void content causes a decrease
in the ILSS of epoxy/CF composites. Therefore, in spite of the good interfacial
interaction between GNPM25 or GNPM15 and TPU-70 HS, as demonstrated in
Chapter Four and in this chapter, the MSCs have weak interfacial adhesion with CF,
resulting in a reduction in the ILSS value compared with the TPU70/CF samples.
The differences in the ILSS values of the TPU70/CF and MSC samples can be
attributed to the differences in void content Vv % and Vf % and to the weak adhesion
between NC matrices and CF, as summarized in Table 5.1 for samples with a
thickness of 4 mm. Hence, it becomes necessary to normalise the value of ILSS of
composites with voids with the ILSS value of composites free from voids, in
accordance with the study conducted by Bowles and Frimpong [334]. The
researchers assumed that all voids in the composites had a spherical shape, and they
derived the relation between the ILSS with voids and the ILSS without voids as
expressed in the following equation:
𝑰𝑳𝑺𝑺𝒗
𝑰𝑳𝑺𝑺𝑭𝑽 = 𝟐/𝟑 ......... (5.4)
𝟔𝑽𝒗
𝟏−[𝟎.𝟕𝟖𝟓∗( ) ]
𝟑.𝟏𝟒∗(𝟏−𝑽𝒇 )

where 𝐼𝐿𝑆𝑆𝐹𝑉 is the ILSS of composites without voids (void-free), 𝐼𝐿𝑆𝑆𝑣 is the ILSS
of composites with voids (as per the experimental results in this study), 𝑉𝑣 is the void
content of the composites and 𝑉𝑓 is the fibre volume fraction.

In accordance with this equation, the normalised value of ILSS, which represents the
ILSS of TPU70/CF and MSCs without voids before and after annealing, are plotted

202
in Figure 5.10. Their normalisation has been carried out according to the high Vf of
samples with 4 mm thickness, which is equal to 47%, and attributed to the TPU70/5
wt.%GNPM5/CF samples (see Table 5.1). The data of the actual value of ILSS and
normalised ILSS of the composites are summarised in Table B. 3 in Appendix B.

50 Non-Annealed
Annealed
Normalised ILSS (MPa)
40

30

20

10

0
CF F F
M5/ 15/C 25/C
70/CF %GNP GNPM GNPM
U t. wt.% wt .%
TP 70/5
w
70/5 70/5
TPU TPU TPU

Figure 5.10: Normalised ILSS of TPU70/CF and MSCs according to void-free


composite equations and Vf = 47% before and after annealing

From this figure, the data of the normalised 𝐼𝐿𝑆𝑆𝐹𝑉 of MSCs exhibits a decrease
from 44  4.6 MPa of TPU70/CF to become 27.5  1 MPa, 27.3  0.3 MPa and 37.2
 0.5 MPa of MSCs with GNPM5, GNPM15 and GNPM25 respectively before
annealing. As with the explanation mentioned previously, this result was ascribed to
the poor interfacial adhesion between NC matrices; GNPM5 and M15 can be
considered a particular reason for this reduction. The current ILSS results can say
compatible with other results reported by Junid [10] despite the difference in the
matrix nature, who found that the addition of A-GNPs to epoxy/CF composites
caused a reduction in the ILSS value from 88 MPa for pristine epoxy/CF samples to
28.8 MPa for MSCs with 6 wt.% A-GNPs. Junid supposed that this reduction was
caused by weak interfacial adhesion between the epoxy NC matrix and poor
distribution of A-GNPs within the epoxy matrix. In addition, Liu [1] demonstrated
that the ILSS of neat epoxy/CF decreased by 9% of its value with the addition of
0.05 wt.% of CNTs; Liu interpreted this as a result of the poor interfacial bonding
between the epoxy/CNT matrix and the CF. Her results and the present results of the
reduction in ILSS due to the addition of nanoparticles to polymer matrices and the

203
effect of poor interfacial bonding between them are in agreement with previously
reported studies [125, 335].

After annealing, the trends of the stress-strain curves of TPU70/CF and MSCs are
almost the same as the results obtained before annealing, as illustrated in Figure 5.9
B. However, the corresponding normalised values of ILSS of the TPU70/CF and
MSCs after annealing are slightly lower than their values before annealing. This
reduction might be connected to the interpretation suggested earlier in this chapter
and in Chapter 4, which proposed that the annealing treatment reduced the
interaction (covalent or HB) between GNPs and the TPU-70 HS matrix, resulting in
restacking of the GNPs and subsequent agglomerations. These agglomerations in
turn reduced the microcrystalline HD, leading to a reduction in phase separation.
From this outcome, it can be deduced that the reduction in ILSS after annealing
might indicate not only poor interfacial adhesion between the NC matrices and the
CF as a result of the void content, but also the reduction in interfacial interaction
between GNPs and TPU-70 HS.

Unlike in the flexural test and tensile test, the ILSS values of TPU70/CF and MSCs
combined with GNPM15 exhibit a slight decline after annealing. This result is
probably due to an increase in the microcrystalline HD of TPU-70 HS and NCs
combined with GNPM15, where the latter acts as a nucleate agent. This
improvement in the microcrystalline HD might reduce the interfacial adhesion
between the TPU-70 HS or NC matrices and CF, which has a strong effect on ILSS
value, in line with earlier reports that the ILSS can provide information about the
interfacial adhesion between fibres and the matrix [114]. Poor interfacial adhesion
could stimulate delamination or fibre-matrix debonding due to the generation of
interlaminar shear failure in the composites, leading to a rupture in maximum load at
a lower value than before annealing. Figure 5.11 and Figure 5.12show the SEM
micrographs of TPU70/CF and MSCs after being subjected to a short beam shear test
(ILSS test). The type of failure mode that occurred in the samples before and after
annealing shows the interlaminar shear failure that caused a fibre-matrix debonding
and delamination for MSCs. However, the TPU70/CF is in fact subjected to flexural
compression failure, which causes a matrix crack and then a breakage in CF not
caused by the interlaminar shear failure, as illustrated in the SEM images in Figure
5.11 A.

204
(A)

Matrix crack

Fibre breakage
20 m

(B)

Delamination

Shear Failure 20 m

Figure 5.11: SEM micrographs of A) TPU70/CF composites and B) TPU70/5


wt.%GNPM5/CF/MSCs after a short beam shear test taken from a side view of the
samples. The right-hand images represent magnified images of the positions circled
in yellow positions in the left-hand image

(C)
Fibre-matrix debonding

Delamination
Shear failure

20 m

205
(D)

Delamination

Shear failure 20 m

Figure 5.12: SEM micrographs of C) TPU70/5 wt.%GNPM15/CF/MSCs and D)


TPU70/5 wt.%GNPM25/CF/MSCs after a short-beam shear test taken from a side
view of the samples. The right-hand images represent magnified images of the
positions circled in yellow positions in the left-hand image
The disparity between the failure modes of the TPU70/CF and MSCs probably refers
to the disparity in the matrix structure. TPU-70 HS possesses high adhesive
properties as a copolymer matrix; this causes it to undergo a smooth failure mode
which causes the curves to rise after reaching a maximum point. The failure begins
as a microcrack inside the matrix, which transfers directly to the fibre due to the
strong interfacial adhesion as demonstrated in the SEM images. Conversely, the
MSCs based on TPU70/GNPs as a matrix indicate the presence of GNPs within the
TPU-70 HS, a reduction in its adhesive properties from one side causes poor
adhesion with the CF; as such, the failure begins as interlaminar shear, which causes
delamination or fibre-matrix debonding, as was also shown in the SEM images.

5.5 Electrical Conductivity

Even though continuous carbon fibres are known for their intrinsic electrical
conductivity, it is expected that the addition of conductive carbon nanofiller can
further enhance this property, especially in the out-of-plane direction. Some
applications of structural composites require high electrical conductivity in the out-
of-plane or through-thickness directions, such as electromagnetic shielding and
aircraft anti-icing or de-icing [336, 337]. Therefore, the integration of CF/polymer
composites with nanofillers that possess relatively high electrical conductivity (such

206
as G, GNPs or CNT) is an established technique for producing MSCs with enhanced
out-of-plane conductivity [111, 146, 148, 338].

In MSCs, the improvement in electrical conductivity depends strongly on several


factors: the concentration of nanofillers, their dispersion in a polymer matrix and
their aspect ratio, the interfacial filler-matrix, filler-filler, and fibre-NC matrix
interactions. Chapter 4 proved that the TPU-70 NCs including 5 wt.% GNPM15 and
GNPM25 reached their percolation threshold at this percentage, hence it is expected
that TPU70-MSCs fabricated with this percentage of GNPM15 and GNPM25 will
fall within the percolation threshold. Therefore, it is now necessary to study the
effect of the size of GNPs on the electrical conductivity of TPU70/CF and MSCs in
both the in-plane and out-of-plane directions.

Figure 5.13 A and B illustrate the electrical conductivity log of TPU70/CF laminates
and MSCs in the in-plane and out-of-plane directions. The corresponding data are
tabulated in Table 5.7 for both directions. In general, it can be clearly seen that the
in-plane conductivity is higher than that of the out-of-plane direction by
approximately three orders of magnitude over the full frequency range. This result
occurs because the direction of carbon fibre reinforcement is in-plane, which forms
an easier conductive pathway than the z-direction (through-thickness). However, the
integration of GNPs causes a significant improvement in electrical conductivity in
the out-of-plane direction and a slight improvement in the in-plane direction. For
instance, the in-plane conductivity of TPU70/CF increased from 422  50 S/m to 449
 15.7 S/m, 439  6 S/m and 559  43 S/m of MSCs combined with GNPM5,
GNPM15 and GNPM25 respectively, as shown in Table 5.7. These results are higher
than those for TPU70/CF by 6%, 4% and 32%, respectively. Despite the significant
improvement caused by integration with GNPM25 compared to other results, it is
still lower than the improvement in out-of-plane conductivity. In brief, the out-of-
plane conductivity is enhanced by 95%, 327% and 286% for MSCs combined with
GNPM5, GNPM15 and GNPM25 respectively in comparison with TPU70/CF. This
rise from 0.02  0.01 S/m of TPU70/CF to become 0.04  0.02 S/m, 0.094  0.06
S/m and 0.085  0.02 S/m of MSCs is recorded for GNPM5, GNPM15 and
GNPM25, respectively.

207
A) In-Plane Direction

Electrical Conductivity (S/m)


103

TPU70/CF
TPU70/5wt. %GNPM5/CF
102 TPU70/5wt. %GNPM15/CF
TPU70/5wt. %GNPM25/CF

100 101 102 103 104 105 106


B) Out-of-Plane Direction
100
Electrical Conductivity (S/m)

10-1

10-2
TPU70/CF
TPU70/5wt. %GNPM5/CF
TPU70/5wt. %GNPM15/CF
TPU70/5wt. %GNPM25/CF

10-3
100 101 102 103 104 105 106
Frequency (Hz)

Figure 5.13: A) In-plane and B) out-of-plane electrical conductivity of TPU70/CF


laminates and MSCs incorporating 5 wt.% GNPs of various size (5, 15 and 25 m
denoted by M5, M15 and M25 respectively)
Table 5.7: Average in-plane and out-of-plane electrical conductivity of TPU70/CF
and MSCs

Sample In-plane conductivity Out-of-plane conductivity


 (S/m)  (S/m)
TPU70/CF 422.3  50.4 0.020  0.01
TPU70/5 wt.%GNPM5/CF 449.0  15.7 0.043  0.02
TPU70/5 wt.%GNPM15/CF 439.0  6.0 0.090  0.06
TPU70/5 wt.%GNPM25/CF 559.0  42.8 0.090  0.02

208
These findings might suggest that the GNPs form more conductive pathways in the
matrix-rich regions, which becomes converse dominant electrical conductivity in in-
plane directions [148]. In contrast, the remarkable improvement in through-thickness
conductivity might be ascribed to the formation of effective networks of GNPs,
especially GNPM15 and M25, leading to the creation of a direct pathway to transport
electrons between the neighbouring carbon-fibre fabric. This finding is compatible
with those of Li et al. [148], who used 5 wt.% GNP to modify epoxy/CF composites.
They reported, respectively, a slight and significant increase in in-plane and out-of-
plane conductivity. The researchers supposed that this increment could be ascribed to
the GNPs’ role in providing a conductive pathway along with all three directions of
the composite structure.

The current results also show agreement with those reported by Kandare et al. [111],
who dispersed GNP and/or silver nanoparticles (SnP/nanowires (SnW)) in epoxy/CF
composites. They reported that the addition of 1 vol% GNP led to a 55% reduction in
the electrical resistivity of the out-of-plane direction. The researchers attributed this
reduction to the formation of graphene sheet networks that create an assistance
pathway to carry electrons between adjacent carbon-fibre fabrics. They also reported
that the combination of GNP and SnP or SnW at the same volume-loading fraction
has approximately 70% more influence on the electrical conductivity of the out-of-
plane direction than GNP alone. They suggested that this improvement might be
attributed to the good physical interaction between GNP and SnP/SnW, leading to a
rise in electrical connectivity.

The geometry and size of nanofillers have an obvious effect on the electrical
conductivity of MSCs in both directions (in-plane and out-of-plane), as demonstrated
in the present work. While both GNPM15 and GNPM25 recorded a high
improvement in out-of-plane conductivity compared with GNPM5, GNPM25 has a
slightly lower value than GNPM15 which conflicts with the expected value. Indeed,
the behaviour of the MSCs combined with GNPM25 is similar to the previous
conductivity behaviour in the out-of-plane direction of NCs combined with
GNPM25, which were discussed in detail in Chapter 4. For the purpose of
comparison, the present results of MSCs integrated with GNPM25 are tabulated with
other earlier results of composite modified by various nanofillers in Table 5.8.

209
Table 5.8: Comparison of in-plane and out-of-plane electrical conductivity between
the current work and previous work on composites modified by various nanofillers

Composites Matrix Fibre Fillers Content  (S/m)  Ref.


% In-plane (S/m)
Out-
of-
plane
CFRP/GNP epoxy carbon GNPs - 6.6 [115]
CFRP/GNP epoxy carbon GNP 5 wt. 101 0.6 [148]
GFRP/CB PE glass CB 10 wt. 2  10 -2 2 [146]
10-2
CFRP/SWCNT epoxy carbon SWCNT 0.25 140×102 0.05× [129]
wt. 102
GFRP/laser- PEEK glass Laser- 1.0 wt.  10-4  10- [123]
SWCNT+PEES SWCNT 5

CFRP/GNPM25 TPU70 carbon GNPM25 5 wt. 559 0.085 Present


work
From Table 5.8, it can be clearly seen that the high in-plane electrical conductivity is
attributed to the MSCs based on the epoxy/CF and incorporated with 0.25 wt.%
SWCNTs compared with the TPU70/CF and MSCs in the current study and in
comparison with other researches. This might be ascribed to the technique that used
to disperse SWCNTs on the surface of CF using a chemical vapour deposition or
might ascribed to high-fibre volume fraction which was unfortunately did not
mention for any one of the above comparison researches. However, the enhancement
in the in-plane electrical conductivity of the TPU70/GNPM25/CF multiscale
composite compared with other tabulated researches in Table 5.8 is quite acceptable,
especially when compared with MSCs based on the epoxy/5 wt.% GNPs/CF and
records an increase by 4 folds. The out-of-plane electrical conductivity of the present
study is lower than that of MSCs based on the epoxy/CF and this might be attributed
as well to the technique that was used to disperse nanofillers on the CF surface or
within epoxy which leads to form a conductive bath.

For example, Qin et al. [115] recorded the highest value of out-of-plane conductivity
(around 6.6 S/m) found in the current study and others research. This result can be
attributed to the techniques that were followed in order to incorporate the GNP in the
epoxy/CF matrix. The researchers used a coating process that involved immersing
the CF in a stable GNP suspension in order to achieve a homogenous dispersion on
the surface of the CF. They found that the rise in out-of-plane conductivity was
around 165%, which they explained by way of the ability of GNP to percolate and

210
generate a consecutive conductive pathway between carbon fibres and GNP
nanoparticles within a composite from one side to another.

5.6 Thermal Conductivity

Atomic lattice vibrations via phonons (pseudo-particles) determine the thermal


conductivity of MSCs. Thermal conductivity is very sensitive to the aspect ratio of
nanofillers, their loading weight, dispersion quality and matrix-nanofiller interfacial
interaction [114, 339]. In the present work, the thermal conductivity of TPU70/CF
and MSCs were calculated using the following equation:

𝒌 = 𝝆𝜶𝑪𝒑 ......... ( 5.5)

where k is the thermal conductivity,  is the density of the sample,  is the thermal
diffusivity and 𝐶𝑝 is the heat capacity at constant pressure. ,  and 𝐶𝑝 are calculated
according to the description mentioned in section 3.11 (Chapter 3). The mapping of
thermal diffusivity of TPU70/CF and MSCs is illustrated in Figure 5.14.

(A) Thermal diffusivity (m-2. s)


(mm)

(mm) Diffusivity (m2/s)

(B)
(mm)

(mm) Diffusivity (m2/s)

211
(C)

(mm)

(mm) Diffusivity (m2/s)

(D)
(mm)

(mm) Diffusivity (m2/s)

Figure 5.14: Pixel versus thermal diffusivity () on the right side and mapping of
thermal diffusivity on the left side of A) TPU70/CF, B) TPU70/5 wt.%GNPM5/CF,
C) TPU70/5 wt.%GNPM15/CF and D) TPU70/5 wt.%GNPM25/CF composites.
The scale bar in this figure is different from one picture to another in accordance
with the structure of the composites and the dispersion of GNPs within the MSCs.
The discrepancy in the colours of the images indicates the quality of dispersion of the
GNPs, where the strongest colour variations represent high thermal heterogeneity
[186]. The calculation of  in this new method has been carried out using non-
contact infrared thermography mapping, and  is linked to the dispersion index (DI).
The DI can be calculated according to the following equations [186]:
𝜶𝒎𝒂𝒙 (𝟏𝟎𝟎𝒑𝒊𝒙𝒆𝒍)−𝜶𝒎𝒊𝒏 (𝟏𝟎𝟎𝒑𝒊𝒙𝒆𝒍)
𝑫𝑰 = ......... (5.6)
𝜶𝒑𝒆𝒂𝒌

where 𝛼𝑚𝑎𝑥 (100𝑝𝑖𝑥𝑒𝑙) and 𝛼𝑚𝑖𝑛 (100𝑝𝑖𝑥𝑒𝑙) represent the maximum and minimum
thermal diffusivity respectively at an arbitrary 100-pixel value, which was considered
as a reference; 𝛼𝑝𝑒𝑎𝑘 represents the thermal diffusivity at the peak of the curve. The

212
value of DI lies between 0 and 1; approaching zero suggests a uniform dispersion
accompanied by a narrow band.

From Figure 5.14, it can be clearly seen that the images of MSCs combined with
GNPM25 are more uniform than those of other samples, which indicates greater
uniformity in the distribution of its thermal diffusivity. The figures situated on the
right-hand side of Figure 5.14 represent the data of thermal diffusivity via the
distribution of each pixel, where thermal diffusivity is located on the x-axis and the
number of pixels is shown on the y-axis. The shape of the curve of thermal
diffusivity distribution can indicate the dispersion quality [186]. A wider and more
variable curve corresponds to non-uniform  distribution, which can refer to non-
uniform GNP dispersion and vice versa. It can be observed from Figure 5.14 that the
MSCs combined with GNPM25 have a narrow curve in comparison with TPU70/CF
and MSCs consolidated with GNPM5 and GNPM15, suggesting that an uniform
dispersion of nanofillers occurs with GNPM25. The calculated values of thermal
conductivity (k), , and Cp of TPU70/CF and MSCs are tabulated in Table 5.9, while
the density of samples with 4 mm thickness was tabulated in Table 5.1.

Table 5.9: Values of thermal diffusivity, thermal conductivity and heat capacity of
TPU70/CF and MSCs

Sample  (mm2/s) Cp (J/g.K) k


(W/m.K)
TPU70/CF 0.080  0.005 1.24  0.02 0.120  0.001
TPU70/5 wt.%GNPM5/CF 0.460  0.002 0.64  0.03 0.380  0.020
TPU70/5 wt.%GNPM15/CF 0.140  0.001 0.89  0.10 0.170  0.020
TPU70/5 wt.%GNPM25/CF 0.123  0.004 0.72  0.07 0.121  0.050
It can clearly be observed from Table 5.9 that the thermal diffusivity () of all MSCs
samples combined with GNPM5, GNPM15 and M25 rises from 0.080  0.005 mm2/s
for TPU70/CF laminates to 0.460  0.002 mm2/s, 0.140  0.001 mm2/s and 0.123 
0.004 mm2/s respectively. This improvement rises by 400 %, 75 % and 54 % for
MSCs combined with GNPM5, GNPM15 and GNPM25 respectively compared to
TPU70/CF composites. The value of thermal conductivity of MSCs combined with
GNPM5 and GNPM25 using equation (5.5) rises by 196% and 32 % respectively
from that of TPU70/CF laminates, while the thermal conductivity of MSCs
incorporated with GNPM25 is marginally unchanged. This discrepancy might be
attributed to the inaccuracy of the measurement of heat capacity due to the small size
of samples that used for testing which are approximately weighing (7-11) mg as
213
described in chapter three ( see sections 3.8.2). However, this technique is focused on
giving an indication to the effect of GNPs sizes on the thermal conductivity of
TPU70/CF laminates and MSCs. In fact, the presence of GNPs regardless of their
size is known to improve the k of polymer matrices due to their high thermal
diffusivity and their ability to form a heat flow network [340, 341].

The highest local values of  and k were recorded for MSCs combined with
GNPM5; this result is attributed to their non-uniform dispersion within MSCs, as
illustrated in Figure 5.14 B. This outcome is compatible with the results reported by
Gresil et al. [186], who revealed and proved this method. The researchers
demonstrated that an epoxy matrix filled with 10 wt.% GNPM25 possesses a wide
and variable curve of  distribution, but that it recorded a higher  value than that of
a narrow curve ascribed to 5 wt.%GNPM25. They explained this by way of the poor
dispersion of 10 wt.% GNPM25 within an epoxy matrix due to the re-stacking and
agglomeration of GNPs with a high loading weight.

In this study, in spite of the weak interaction between GNPM5 and a TPU-70 HS
matrix from one side and CF from the others as demonstrated previously in this
chapter (see sections 5.3, 5.4.1 and 5.4.2), these particles recorded the highest value
of k. However, this high local thermal conductivity of GNPM5-MSCs is not uniform
across the whole sample. This unexpected finding might be attributed to the high
fibre volume fraction of MSCs consolidated with GNPM5 (475%), leading to a
reduction in the thickness of matrix interlayers and a subsequent improvement in the
heat flow network. Imran and Shivakumar [147] found that the impregnation of CF
fabric with a dispersion of GNPM25 and epoxy using a hand lay-up process followed
by compression moulding caused a slight improvement in the thermal conductivity of
epoxy/CF by 6%. Conversely, the researchers found that the addition of 1
wt.%GNPM25 to an epoxy matrix doubled the thermal conductivity of the epoxy.
They attributed this discrepancy in improvement between the NCs and MSCs to the
effect of the CF volume fraction when the Vf of epoxy/CF is higher than that of the
CF/GNPM25/epoxy, measured at 57% and 51% respectively. This proposal is in
agreement with the present suggestion concerning the effect of V f on increased
improvement in the thermal conductivity of GNPM5-MSCs.

214
In line with the previous discussion regarding electrical conductivity, Li et al. [148]
found that modifying the epoxy/CF composite by 2 wt.% GNP and 5 wt.% GNP
caused improvements in the out-of-plane thermal conductivity by 8% and 50%
respectively. They attributed this enhancement to the high surface area and flat shape
of GNP, which allows considerable areal contact between adjacent platelets within
the epoxy matrix and reduces the phonon-scattering that occurs owing to
incompatibilities in the thermal conductivity between the insulating epoxy and
thermally conductive GNPs. Similarly, Kandare et al. [111] found that the
consolidation of 1 vol.% GNP and SnP/SnW as hybrid nanofillers caused an
improvement in through-thickness thermal conductivity of epoxy/CF by 40%
compared with 9% of 1 vol.% GNP only.

5.7 Summary

The main purpose of fabricating multiscale composites consisting of TPU/GNPs/NCs


and reinforced with CF is to improve their through-thickness mechanical, thermal
and electrical properties. Choosing three different sizes of GNPs (5, 15 and 25) m
in average diameter and studying their effect on thermoplastic polymers such as TPU
reinforced with CF fabric situates this work as innovative in the field of MSCs.
Despite many difficulties that arose during this work from the fabrication of MSCs to
testing, it is noteworthy that this investigation has fulfilled its targets, which can be
summarised in the below points:

• The size of GNPs has been found to have a significant effect on the mechanical
and thermal properties of MSCs compared to TPU70/CF. For example, in the
DMTA test, the E values of MSCs combined with GNPM15 and GNPM25
were higher than those of TPU70/CF over the entire temperature range.
Meanwhile, MSCs combined with GNPM5 exhibited a value lower than
TPU70/CF across the whole temperature range before and after annealing. The
TgHS has been found to be significantly improved by the addition of GNPs,
regardless of their size in comparison with TPU70/CF. The impact of annealing
was found to be a significant reduction in E’ and TgHS of TPU70/CF and
MSCs, which was attributed to the effect of annealing on the formation of
microcrystalline HD and then on the phase separation.

215
• In the tensile and flexural results, the addition of GNPs to TPU70/CF brought
about a spectacular improvement in the FM of MSCs combined with GNPM15
and M25 in comparison with TPU70/CF, as well as a slight improvement in .
In addition, the annealing treatment yielded a reduction in the flexural and
tensile properties of MSCs consolidated with GNPM25, while TPU70/CF and
MSCs incorporating GNPM5 and M15 showed an impressive enhancement.

• For the ILSS test, the results were unexpected because all the MSCs exhibited
a reduction in ILSS compared to TPU70/CF before and after annealing. The
maximum reduction was caused by GNPM5 due to its weak interaction with
the TPU-70 HS matrix on one side and its NCs with CF on the others, resulting
in premature failure due to delamination and fibre-matrix debonding.

• In general, the MSCs showed a slight increase in in-plane electrical


conductivity and a considerable improvement in out-of-plane conductivity
compared to TPU70/CF.

• The through-thickness thermal diffusivity improved for TPU70/CF after the


addition of GNPs, regardless of their size. An unexpected outcome was that the
maximum improvement was attributed to MSCs combined with GNPM5.

Concluding from the above brief points concerning the effect of GNP size on the
mechanical, electrical and thermal properties of TPU70/CF laminates and MSCs,
GNPM25-MSCs can be nominated as the best candidate for the formation of
TPU70/CF-MSCs, even though the particles exhibited a small failure in enhancing
the thermal conductivity and ILSS. Therefore, GNPM25 particles are chosen to form
MSCs with different percentages lower than 5 wt.%. As previous works have
demonstrated, levels of up to 5 wt.% of nanofiller within MSCs led to a deterioration
in mechanical properties; this work has aimed to avoid sacrificing mechanical
properties for improvement in electrical and thermal properties.

On the other hand, the annealing treatment will not be applied in the next chapter
since this technique improved the properties of neat TPU-70 HS and reduced those of
NCs, TPU70/CF laminate and MSCs especially those combined with GNPM25.
Therefore, using this technique is not necessary in the aim of investigating the effect
of GNPs on the damage resistance and damage tolerance of composites laminates.

216
Chapter 6: The Effect of GNPM25 on the
Delamination and Damage Tolerance of
TPU70/CF Composites

6.1 Introduction

Interlaminar failure or delamination is well known as a crucial problem in fibre


composites systems. Composites in structural applications are exposed to limiting
factors attributed to the delamination failure associated with matrix cracking or
fibre/matrix debonding. As a result, catastrophic failure of composite structures can
occur due to reductions in strength and stiffness as a consequence of the initiation
and propagation of delamination [126]. Delamination as a form of internal damage
may lead to a large decline in the residual compressive strength of composite
structures when subjected to low-energy impact [342]. Therefore, most efforts have
focused on improving the damage tolerance and delamination resistance of
composites by improving matrix toughness. Recently, the use of thermoplastic
polymers to reinforce carbon fibre fabric has attracted the attention of many
researchers due to their potential advantages in comparison with thermoset polymers.
The main advantages include high damage tolerance, shorter manufacturing time and
superior toughness. Consequently, including GNPs in thermoplastic matrices might
reduce their damage tolerance and delamination resistance, especially at high
nanoparticle loading, due to the occurrence of agglomeration. It is well known that
graphene and its derivative tend to restack together due to the attraction caused by
Van der Waal forces between graphene layers at high loading contents, which
constitute a major problem in the field of nanocomposites and MSCs. In order to
avoid this problem and because of the decrease in the mechanical properties of MSCs
at high GNP loading, loadings of 0.5 wt.%, 2.5 wt.% and 5 wt.% of GNPM25
combined with TPU-70 HS and reinforced with CF were selected for this work.

In this chapter, theoretical micromechanics models will be used to predict the elastic
modulus and aspect ratio of NCs combined with GNPM25, which was nominated in
the previous chapter as the optimal size of GNP to be consolidated with neat TPU-70

217
HS and TPU70/CF composites. The new MSCs incorporated with different loadings
of GNPM25 will be characterised in terms of their delamination and impact
resistance and damage tolerance after being subjected to a mode-I interlaminar
fracture toughness (ILFT) test, low-velocity impact tests and a compression-after-
impact test (CAI). The SEM technique will be applied to investigate the fracture
surface after the mode-I ILFT and low-velocity impact tests in order to identify the
type of failure.

6.2 Micromechanical Modelling Analysis

Over several decades, the development of theoretical models has played a significant
role in predicting the properties of composite materials based on the properties of
their constituents. This section intends to predict the elastic modulus of
TPU70/GNPM25/NCs at different weight percentage loadings (0.5, 2.5 and 5). The
objective of this study is to predict the elastic modulus of NCs depending on the
extrapolation when the E of GNPs is assumed equal to 15, 30, 36.5, 50, 100, 150 and
250 GPa instead of using the common value of 36.5 GPa, as illustrated in Chapter 5
(see section 5.4.1.1).

6.2.1 Halpin-Tsai Model

The prediction of the elastic modulus of nanofiller-reinforced composites can be


estimated using different models depending on the geometry, orientation, weight
percentage and elastic properties of matrices and fillers. Halpin-Tsai equations are
considered one of the simplest and most common models that can be adopted to
predict the tensile moduli of polymer composites using the aspect ratio, volume
fraction and elastic modulus of fillers and matrices. The Halpin-Tsai model [343] can
predict the tensile modulus of nanocomposites that consist of fillers oriented in both
longitudinal and transverse directions [128, 258, 344, 345]. The equations below
show the original forms of the Halpin-Tsai (H-T) model:

𝑬𝒄 𝟏+𝑽𝒇
= ....... ( 6.1)
𝑬𝒎 𝟏−𝑽𝒇

(𝑬 /𝑬 )−𝟏
 = (𝑬𝒇 /𝑬𝒎 )+ ....... ( 6.2)
𝒇 𝒎

where Ef, Em and Ec are the elastic modulus of fillers, matrix and composites,
respectively; Vf is the volume fraction of fillers; and  is the shape factor that
218
depends directly on the filler geometry and loading direction. In general,  = 2(L/d)
for fibres, where L is the fibre length and d is the fibre diameter, and  = 2(L/t) for
disk-like platelets, which are adopted in this study, where L is the GNP length and t
is the GNP thickness. The shape factor , which relates to the aspect ratio of
reinforcement, has provided good agreement with the longitudinal modulus;
conversely, the transverse modulus was found to be comparatively insensitive to
aspect ratio [128]. It should be noted that as →0, the Halpin-Tsai equation (4.1)
reduces to a lower bound (iso-stress) of the rule of mixture model, while when →,
Equation 4.1 reduce to the upper bound (iso-strain) of the rule of mixture model.

𝑼𝒑𝒑𝒆𝒓 𝒃𝒐𝒖𝒏𝒅 𝑬 𝒄 = 𝑽𝒇 𝑬 𝒇 + 𝑽𝒎 𝑬 𝒎 ....... ( 6.3)


𝟏 𝑽 𝑽
𝑳𝒐𝒘𝒆𝒓 𝒃𝒐𝒖𝒏𝒅 = 𝑬𝒇 + 𝑬𝒎 ....... ( 6.4)
𝑬𝒄 𝒇 𝒎

For composites with unidirectional or discontinuous fillers, the Halpin-Tsai models


can predict the composite elastic modulus in the longitudinal (EL) and transverse (ET)
directions according to the equations below:
𝟏+ 𝑽
𝑬𝑳 = 𝑬𝒎 ( 𝟏− 𝑳𝑽 𝒇 ) ....... ( 6.5)
𝑳 𝒇

𝟏+𝟐 𝑽
𝑬𝑻 = 𝑬𝒎 ( 𝟏− 𝑻𝑽 𝒇 ) ....... ( 6.6)
𝑻 𝒇

where the parameters L and T can be expressed in the following equations:

(𝑬𝒇 /𝑬𝒎 )−𝟏


𝑳 = (𝑬𝒇 /𝑬𝒎 )+
....... ( 6.7)

(𝑬 /𝑬 )−𝟏
𝑻 = (𝑬𝒇 /𝑬𝒎 )+𝟐 ....... ( 6.8)
𝒇 𝒎

To follow the Halpin-Tsai models, the below assumptions should be applied to the
TPU70/GNPM25/NCs system:

i. Both GNPs and TPU-70 HS are isotropic and firmly bonded to each other;

ii. The GNPs are asymmetrical, identical in shape and size, homogeneously
dispersed and aligned;

iii. The GNPs are fully exfoliated and perfectly oriented.

219
Indeed, it can be clearly observed that these assumptions will result in a disparity
between the predicted Halpin-Tsai elastic moduli and the real behaviour of the
complex NCs system (TPU70/GNPM25). These disparities will be discussed and
explained later in this chapter. To study the effects of filler orientation within the
matrix, the Halpin-Tsai equations have been modified according to the possible filler
orientation, which may be 2D or completely 3D and which may be randomly
oriented [346-349]. The modified Halpin-Tsai (MH-T) equations for 2D and 3D
random orientations can be written as shown below [346-349]:
𝟑 𝟓
𝟐𝑫 𝒓𝒂𝒏𝒅𝒐𝒎𝒍𝒚 𝒐𝒓𝒊𝒆𝒏𝒕𝒆𝒅 𝒇𝒊𝒍𝒍𝒆𝒓 𝑬𝒄 = 𝟖 𝑬𝑳 + 𝟖 𝑬𝑻 ....... ( 6.9)
𝟏 𝟒
𝟑𝑫 𝒓𝒂𝒏𝒅𝒐𝒎𝒍𝒚 𝒐𝒓𝒊𝒆𝒏𝒕𝒆𝒅 𝒇𝒊𝒍𝒍𝒆𝒓 𝑬𝒄 = 𝟓 𝑬𝑳 + 𝟓 𝑬𝑻 ....... ( 6.10)

Equations (6.9) and (6.10) can be also written in different way as illustrated in
equations (6.11) and (6.12 ) according to van Es et al. [350] using laminate theory
[128, 350].
𝟐𝑫 𝒓𝒂𝒏𝒅𝒐𝒎𝒍𝒚 𝒐𝒓𝒊𝒆𝒏𝒕𝒆𝒅 𝒇𝒊𝒍𝒍𝒆𝒓 𝑬𝒄 = 𝟎. 𝟏𝟖𝟒 𝑬𝑳 + 𝟎. 𝟖𝟏𝟔 𝑬𝑻 …… (6.11)

𝟑𝑫 𝒓𝒂𝒏𝒅𝒐𝒎𝒍𝒚 𝒐𝒓𝒊𝒆𝒏𝒕𝒆𝒅 𝒇𝒊𝒍𝒍𝒆𝒓 𝑬𝒄 = 𝟎. 𝟒𝟗 𝑬𝑳 + 𝟎. 𝟓𝟏 𝑬𝑻 ...... (6.12)

In this study, Em of TPU-70 HS is 0.2 GPa, as measured previously (see section


4.3.6.1); the aspect ratio (L/t) of GNPM25 is 3125 when L and t represent the length
and average thickness of GNP, which were taken as 25m and 8 nm respectively
according to the supplier specifications [150]. Finally,  was taken as 2(L/t); in this
case, it is equal to 6250.

6.2.1.1 Effect of Variation in Filler Modulus and Aspect Ratio


Figure 6.1 illustrates the comparison between predicted and actual measured data of
the elastic modulus of TPU70/GNPs/NCs. From this figure a deviation between the
predicted and experimental values of the elastic modulus of nanocomposites can be
clearly observed. The predicted theoretical modulus of TPU70-NCs increases with
increasing Ef and Vf. Specifically, it can be observed that the experimental results fit
at Ef =100 GPa at loading <1 wt.% and lie between 15 and 30 GPa at a loading of > 1
wt.%. This means that the assumption of considering the Ef of GNPs equals to 36.5
GPa as depended in chapter five (see section 5.4.1.1) does not fit with experimental
results at loading higher or lower than 1 wt. % GNPs. The reduction in Ef at high
loading could be a result of the assumptions of homogeneity in dispersion, alignment

220
and identical shape and size of the GNPs, while there is likely to be some
agglomeration or poor dispersion in the real behaviour of NCs [344, 351].

30

Experimental Modulus
25 Halpin-Tsai Modulus Ef=250GPa

20 Em=0.2GPa
EC/Em

= Ef=150GPa

15
Ef=100GPa

10 Ef=50GPa
Ef=36.5GPa
5 Ef=30GPa
Ef=15 GPa

0
0 1 2 3 4 5 6
GNP Content (wt. %)

Figure 6.1: Halpin-Tsai model showing the effect of Ef on the tensile modulus of
TPU-70 HS nanocomposites
In earlier work, Ahmad et al. [352] reported the effective elastic modulus of
GNPM15 reinforced with PP polymer and prepared by melt compounding and then
injection moulding to be around 100 GPa. This result is compatible with the present
results, in spite of the differences in GNP aspect ratio and processing methods. In
addition, Albozahid [46] found that the MH-T in 2D randomly oriented filler fitted
with the experimental results at 50 GPa at a loading of < 3 wt.% GNPM15 with
TPU-70 HS and prepared by in situ polymerisation followed by injection moulding.
The difference between this study and the latter may be related to the high aspect
ratio of GNPM25, which tends to increase the elastic modulus of NCs, as
demonstrated in Chapter 4 of this study.

As a result of these findings, an Ef of 100 GPa was chosen to generate further


modelling to predict the elastic modulus of TPU70/GNPM25 nanocomposite; this
value will be extrapolated by varying the GNP aspect ratio (L/t). Figure 6.2 shows
the ratio of elastic modulus of TPU70/GNPs/NCs to neat TPU-70 HS as a function of
GNP loading. The value of Em is 0.2 GPa, as in the previous model. The lines
marked U and L represent the calculations of predicted elastic modulus using the
upper and lower bounds (series and parallel) of the rule of mixture models described

221
in equations (6.3) and (6.4) respectively. The experimental results are marked with
red circles, indicating their standard deviation. The calculation of the Halpin-Tsai (H-
T) model has been performed using equations (6.1) and (6.2). The aspect ratio varies
from 30 to 1000.

U
12 Experimental Modulus
Halpin-Tsai Modulus L/t=1000
Lower bound
10
Upper bound
L/t=500

8
EC/Em

Em=0.2GPa
L/t=200
6 Ef= 100GPa

L/t=100
4
L/t=50
L/t=30
2
L
0
0 1 2 3 4 5 6
GNP Content (wt. %)

Figure 6.2: Calculation of Halpin-Tsai model showing the effect of loading and
aspect ratio of GNPs on the tensile modulus of TPU70-NCs incorporated with 0.5
wt.%, 2.5 wt.% and 5wt.% GNPM25. The letters U and L indicate the upper and
lower bound respectively
The assumptions in this model are similar to those mentioned previously in section
6.2.1; As these assumptions were built on full exfoliation and uniform dispersion of
GNPs, any discrepancy with the experimental data can be attributed to
agglomeration, poor dispersion, rolling up or folding of the GNP platelets [46, 208,
219]. As such, these types of models, albeit advantageous, can only give a general
description of the influence of filler loading, elastic modulus and aspect ratio on the
tensile modulus of NCs [344].

It can be clearly seen in the above figure that the elastic modulus of TPU70-NCs
increases with increasing aspect ratio and GNP loading. As such, the results of the H-
T theoretical models lie between the upper and lower bounds of the rule of mixture
models. At low loading of GNPs (< 1 wt.%) the experimental data fits with the line
denoted as L/t = 1000 and then drops at a loading of > 1 wt.% to lie exactly between
L/t = 100 and L/t = 50. At higher loading, the reduction in the aspect ratio of GNPs is
in all probability due to the formation of agglomerations or restacking of GNP flakes

222
[208, 219, 353]. A previous study of the H-T model [219] reported that the
experimental results of PA6/NCs fitted on the line delineating L/t = 2000 at a loading
of < 1 wt.% GNPM25. The difference with the current results could be attributed to
the difference in the preparation of the NCs, or is likely to refer to the poor
interaction between GNPs and the TPU-70 HS matrix in comparison with PA6.
Conversely, another study [208] demonstrated that the experimental results of
PET/NCs lie on the line defining the aspect ratio between 200 and 500 at a loading of
< 2 wt.% GNPM15. In this case, the discrepancy with the current results might refer
to the high surface area of GNPM25, which might produce better interaction with the
TPU-70 HS matrix.

6.2.1.2 Prediction of Tensile Modulus of TPU70/GNP Nanocomposites


Figure 6.3 illustrates a comparison between the original H-T model and MH-T
models in 2D and 3D random orientations using equations (6.1), (6.2), (6.9) and
(6.10), respectively. This study depends on using equations (6.9) and (6.10) as the
more common equations that used to predict the elastic modulus of NCs in 2D and
3D random orientations than equations (6.11) and (6.12) as reported by
Alshammari’s study [208] and Albozahid’s study [46].

12 Experimental Modulus
Original Halpin-Tsai Modulus
MH-T-2D randomly oriented
10 MH-T-3D randomly oriented
EC/Em

8
Em=0.2 GPa
Ef= 100 GPa
6
L/t=3125

0
0 1 2 3 4 5 6
GNP Content (wt. %)

Figure 6.3: Tensile modulus of elasticity of TPU70/GNPM25/NCs: Comparison


between Halpin-Tsai model, 2D MH-T and 3D MH-T models with experimental data

223
The comparison was performed using Em and Ef as 0.2 and 100 GPa respectively,
while the aspect ratio of GNPM25 was taken as an actual value of 3125. It can be
clearly seen that the experimental data gives a better correlation with the H-T model
at a GNP loading of less than 1 wt.%, which is deviated upon high loading.
Conversely, the MH-T for 2D or 3D randomly oriented filler does not fit with the
experimental data at low or high loading, while the experimental data at high loading
lies between them. It should be mentioned that the assumptions which were followed
for all models in this calculation were the same as the assumptions which were built
into the Halpin-Tsai model.

This finding has good agreement with previous work that used H-T models and/or
MH-T models [102, 227, 348, 351, 354]. For instance, King [348] revealed a better
match between the experimental data of epoxy/GNPs nanocomposite and a 2D
randomly oriented MH-T model more than with a 3D randomly oriented model. The
researcher attributed this difference to the high-shear mixing that was used during the
fabrication process.

Kalaitzidou et al. [351] found a good agreement between experimental data and an
H-T model of GNP/PP prepared by extrusion and injection moulding at low loading,
but found that the results were over-predicted at high loading. Shookrieh et al. [102]
demonstrated that the results of tensile modulus of epoxy/GNPs nanocomposite
predicted by H-T models fitted with measured results at a loading of < 0.25 wt.% and
deviated at high loading. They attributed this deviation to the non-uniform dispersion
that conflicted with the assumptions of uniform dispersion followed in these
theoretical models, neglecting the effect of real dispersion features such as the
misalignment of GNP platelets within the epoxy matrix.

Bandla and Hanan [227] found poor matches between the theoretical tensile modulus
of PET/GNPs estimated using an H-T model and measured values. The researchers
attributed these results to the H-T models’ assumption relating to the homogeneous
alignment of GNP platelets, when misalignment or random distribution is more
likely to be the case. Their results agree with those reported by Duguay et al. [354],
which demonstrated poor compatibility between the experimental results and H-T
model for tensile modulus. They attributed these findings to the high theoretical
value of aspect ratio that was used in the model. In fact, during melt compounding,
there was an agglomeration or distortion of GNP platelets resulting from buckling or

224
rolling up, which can lead to a reduction in the aspect ratio of GNPs compared with
the theoretical value, which was built on full exfoliation.

6.3 Mechanical Testing of Multiscale Composites

The average properties of all samples used for mode-I ILFT tests, low-velocity
impact tests, and compression-after-impact tests are listed in the table below:

Table 6.1: Average properties of TPU70/CF and multiscale composites


Carbon-
Laminate Density of fibre Matrix Voids
volume volume
Specimens thickness composites volume
fractions fraction
(mm) (g/cm3), 𝝆𝒄 fractions (Vm), (%) (Vv)%
(Vf), (%)
TPU70/CF 3.7  0.1 1.37  0.01 43.1  0.2 56.7  0.3 0.3  0.1
TPU70/0.5 wt.%GNP/CF 3.6  0.1 1.38  0.01 43.4  1.1 56.0  1.2 0.6  0.2
TPU70/2.5 wt.%GNP/CF 3.7  0.0 1.36  0.02 44.0  2.1 54.0  3.6 1.8  0.2
TPU70/5 wt.%GNP/CF 3.7  0.1 1.37  0.02 45.4  0.9 51.6  0.4 3.7  0.1
From this table, it can be concluded that the differences in Vf of TPU70/CF
laminates and MSCs is not high enough to be a cause for concern or to require
normalisation, as was the case in the mechanical tests in Chapter 5. Furthermore, as
found previously regarding the difference in the Vf of composites, if the difference
was  5%, there was no effect on the interlaminar fracture toughness or critical strain
energy release rate (GIC) [355]. Therefore, due to the lower variation in the Vf of
TPU70/CF and MSCs and as a result of the type of tests that have been done in this
chapter, where the matrix is more dominant than the fibres, this study will use actual
testing values without normalisation according to the high Vf.

6.3.1 Mode-I Interlaminar Fracture Toughness Tests

The relation between the cubic root of compliance (C1/3) and the crack length, a, of
TPU70/CF and MSCs is plotted in Figure 6.4. The compliance, C, represents the
ratio between the displacement, , and the applied load, P [201]. As seen in this
graph, the gradient of the TPU70/CF composite reduces with the presence of GNPs,
and this reduction grows as the weight percentage of the GNPs increases. This result
is because of the increase in stiffness of MSCs with increasing GNP content [10].

225
The correction factor for the crack-tip rotation (|∆|), which is used to calculate the
mode-I strain energy release rate, GI, can be determined by the linear regression of
C1/3 as a function of crack length. That is, |∆| represents the negative interception of
the best-fitting line through the test data and the crack length x-axis [135].

0.9
(A) 0.9
(B)

0.8 0.8

test values test values


C1/3 (mm/N)1/3

0.7
Linear fit
0.7 Linear fit

0.6 0.6

0.5 0.5
y=0.00827x+0.04853 y=0.00816x+0.08166
0.4
2
R =0.99417 0.4 R2=0.99236

0.3 0.3
30 40 50 60 70 80 90 30 40 50 60 70 80 90

(C) 0.9
(D)
0.9

0.8 0.8

test values
C1/3 (mm/N)1/3

0.7 0.7 Linear fit


test values
Linear fit
0.6 0.6

0.5 0.5
y=0.00788x+0.10058 y=0.00717x+0.16526
2
R =0.98713 R2=0.98951
0.4 0.4

0.3 0.3
30 40 50 60 70 80 90 30 40 50 60 70 80 90
Crack length,a (mm)
Crack length,a (mm)

Figure 6.4: Representation of cube-root compliance C1/3 versus crack length, a, of


A) TPU70/CF; B) MSCs incorporated with 0.5 wt.% GNPs; C) MSCs incorporated
with 2.5 wt.% GNPs; and D) MSCs incorporated with 5 wt.% GNPs
Figure 6.5 illustrates the typical load-displacement curves of all four samples. As can
be clearly seen from this figure, all samples experience an initial stable linear rise of
load-displacement curves followed by a stick-slip or unstable declining region during
crack propagation. In fact, this type of behaviour accompanies the fabric
reinforcement, as described by Kuwata and Hogg [356]. The researchers applied this
test to composites reinforced with three different types of weave fabric:
unidirectional, satin and plain. Afterwards, they found that the shape of the load-

226
displacement curve depends directly on the type of weave fabric. This variation in
the load-displacement curve was associated with the mechanisms of crack initiation
and propagation. It was demonstrated that the initiation of the crack for composites is
similar for all three types of fabric, but that they differ in terms of crack propagation
[356]. Therefore, the crack propagation behaviour of TPU70/CF and MSCs
reinforced with 5 harness satin (5H) CF fabric is characteristic of stick-slip
behaviour, in that it jumps from one point to another. This behaviour is the same as
that which was reported by Kuwata and Hogg [356], in spite of the differences in
composite matrices.

200
TPU70/CF
175 TPU70/0.5wt.%GNP/CF
TPU70/2.5wt.%GNP/CF
150 TPU70/5wt.%GNP/CF

125
Load (N)

100

75

50

25

0
0 10 20 30 40 50
Displacement (mm)

Figure 6.5: Typical load-displacement curves of Mode-I test (DCB) of TPU70/CF


laminates and MSCs incorporated with 0.5, 2.5, and 5 wt. % GNPs with 25 m
average particle sizes
The main reason for stick-slip crack propagation is the effect of transverse fibre
bundles of 5HS weave fabric, which caused the crack to become deflected, bowed or
branched. Other researchers have attributed this behaviour of the composite or MSCs
to differences in material properties, such as fibre-rich regions in the in-plane
direction or matrix-rich regions, voids in the composites, misalignments, or indeed
fibre bridges or bundles [304, 357].

227
It can be noticed from Figure 6.5 that the MSCs incorporating a lower amount of
GNPs show a high load-carrying capacity before crack growth, as well as a long
displacement compared to the reference composites and other MSCs. These
characteristics can be attributed to the improvement in the composite toughness,
suggesting higher resistance to crack initiation and crack growth. As the high content
of GNPs causes a significant reduction in load-carrying capacity, the crack
propagates fast and catastrophically. For example, the maximum load that TPU70/CF
laminate can sustain before propagating in delamination is 171 N at a displacement
of around 7 mm, which increases to reach 179 N and 9 mm displacement by
consolidating 0.5 wt.% GNP. The MSCs which were combined with 2.5 wt.% GNPs
show almost the same maximum load of TPU70/CF with a slight increase in
displacement, while the MSCs with high GNP content (5 wt.%) exhibit a
considerable reduction in load and displacement of around 120 N and 5 mm
respectively. These observations might be attributed to the poor adhesion between
TPU70-NCs and CF, especially at high content when the agglomeration might take
place [335, 358, 359]. On the other hand, the void contents increase as the weight
percentage of GNPs increases, recording a high value at 5 wt. %. These void contents
(see Table 6.1) and the agglomeration of GNPs in matrix-rich regions can together
act as points of stress concentration in the composite structure, resulting in a
reduction in sustained load [316, 317].

The critical strain energy release rate (GIC) of crack initiation and propagation are
determined using load-displacement curves according to the ASTM D5528 and
following modified beam theory (MBT) (described in Chapter 3, section 3.14.4). The
GIC of crack initiation has three different values: the first corresponds to the point
when the load deviates from linearity in the load-displacement curves (GIC-NL). This
value assumes that a crack starts to grow from the insert in the interior sample [135,
201]. The point when the delamination can be observed on the either edge (GIC-VIS)
represents the second value. The final value is at the point of maximum load or at the
point when the compliance has increased by 5% (GIC-Max). For crack propagation, the
average value of GIC-Prop is evaluated in the crack propagation region after the GIC-Max
value. The results of all values of GIC of TPU70/CF and MSCs are illustrated in
Figure 6.6 and tabulated in Table 6.2. As seen from this figure and table, TPU70/CF
laminates and MSCs experience all three values of GIC of crack initiation, which can

228
indicate a tough matrix. It is known that a brittle matrix has almost the same value of
GIC-NL and GIC-VIS, whereas these points clearly identify tough matrices [135, 201].

Table 6.2: Results of Mode-I fracture-toughness test of TPU70/CF and MSCs using
MBT, including the three values of GIC for crack initiation (GIC-NL, GIC-VIS and GIC-
Max) and average GIC-Prop during crack propagation

Sample GIC-NL GIC-VIS GIC-Max Average GIC-


(kJ/m2) (kJ/m2) (kJ/m2) 2
Prop (kJ/m )
TPU70/CF 1.3  0.2 1.7  0.3 3.09  0.37 2.70  0.40
TPU70/0.5 wt.% GNP/CF 1.6  0.4 2.5  0.4 3.15  0.40 2.99  0.40
TPU70/2.5 wt.% GNP/CF 1.3  0.4 2.0  0.2 2.70  0.18 2.32  0.10
TPU70/5 wt.% GNP/CF 1.1  0.3 1.3  0.1 1.60  0.10 1.20  0.10

5.5
TPU70/CF
5.0 TPU70/0.5wt.%GNP/CF
4.5 TPU70/2.5wt.%GNP/CF
TPU70/5wt.%GNP/CF
4.0
GIC (kJ/m2)

3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
30 40 50 60 70 80 90
Crack length,a (mm)

Figure 6.6: Typical R-curves (critical strain energy release rate, GIC) of TPU70/CF
and MSCs versus crack length
As mentioned previously, adding a lower weight percentage of GNPs (0.5 wt.%)
causes a slight increase in the load-carrying capacity in comparison with the
reference sample of TPU70/CF. This improvement in turn causes a moderate
enhancement in the initiation and propagation of GIC. For example, the GIC-NL, GIC-
VIS, GIC-Max and GIC-Prop of MSCs incorporated with 0.5 wt. % GNPs increase by 23

229
%, 44 %, 10 % and 10 % respectively in comparison with TPU70/CF composites. It
can be clearly noted that the greater improvement obtained by incorporating 0.5
wt.% GNPs occurs at GIC-NL and GIC-VIS, suggesting that a higher energy is required
to initiate a delamination crack. The requirement of this high energy is due to the
presence of GNPs that might pin and deflect the crack; as a result, high fracture
toughness is attained [304]. Conversely, the 2.5 wt.% incorporation produces a slight
increase in GIC-NL and GIC-VIS by 2 % and 14 %, and the GIC-Max and GIC-Prop reduce by
12% and 14%, respectively. All values of MSCs incorporating 5 wt.% GNPs show a
reduction in comparison with TPU70/CF. A significant reduction of GIC-Max and
average GIC-Prop is exhibited, where they reduce by 50% and 56% from 3.0 kJ/m2 and
2.7 kJ/m2 of TPU70/CF to 1.6 kJ/m2 and 1.2 kJ/m2 respectively. The previous
explanation for this reduction, which relates to poor adhesion between NC matrices
at high GNP loading and CF, is also applicable here. In spite of good interfacial
interaction between GNPM25 and TPU-70 HS, as demonstrated earlier (see Chapter
4), the higher loading of GNPs results in an agglomeration that works as a flaw or
defect, resulting in a point of stress concentration. Therefore, the presence of stress
concentration due to the agglomeration or intralaminar delamination due to poor
adhesion could be the major causes of the considerable reduction in fracture
toughness of MSCs incorporated with 2.5 wt.% GNPs and 5 wt.% GNPs
respectively.

To further understand the failure mechanism, a SEM technique was used to


investigate the fracture surface at points near and far from the inserted film. Figure
6.7 to Figure 6.10 show the fracture regions of TPU70/CF, MSCs incorporated with
(0.5, 2.5 and 5) wt.% GNPs, respectively. The most important observation that can
be made from these figures is that the fracture surface of TPU70/CF and MSCs
incorporated with 0.5 wt.% GNPs is smooth, while it becomes rougher with the
increase in GNP content. The roughness of this surface can indicate the GNP
agglomeration regions (mainly in the matrix-rich region), which can be seen in
Figure 6.9 and Figure 6.10 respectively. The rougher surface is attributed to the
MSCs with higher GNP content, suggesting a larger size and amount of
agglomeration is formed at high GNP loading. The presence of substantial
agglomeration could facilitate crack propagation without pinning or bifurcation,
leading to a reduction in the interlaminar fracture toughness (ILFT).

230
A) Fracture near the inserted film of TPU70/CF

10 m

B) fracture surface away from the inserted film of TPU70/CF

Figure 6.7: SEM images of TPU70/CF composites at fracture surface in Mode-I


fracture toughness testing A) near the inserted film and B) far from the inserted film.
The yellow lines represent the hackles, while the blue lines represent the river
patterns. The yellow rectangular area indicates the magnified region

231
A) Fracture near the inserted film of 0.5wt.%GNP-MSCs

10 m

B) fracture surface away from the inserted film of 0.5wt.%GNP-MSCs

Figure 6.8: SEM images of MSCs combined with 0.5 wt.% GNPs at fracture surface
of Mode-I fracture toughness testing A) near the inserted film and B) far from the
inserted film. The yellow lines represent the hackles, while the blue lines represent
the river patterns. The yellow rectangular area indicates the magnified region.

232
A) Fracture near the inserted film of 2.5wt.%GNP-MSCs

10 m

B) fracture surface away from the inserted film of 2.5wt.%GNP-MSCs

Figure 6.9: SEM images of MSCs combined with 2.5 wt.% GNPs at fracture surface
of Mode-I fracture toughness testing A) near the inserted film and B) far from the
inserted film. The red lines represent the agglomeration regions of GNPs

233
A) Fracture near the inserted film of 5wt.%GNP-MSCs

10 m

B) fracture surface away from the inserted film of 5wt.%GNP-MSCs

Figure 6.10: SEM images of MSCs incorporated with 5 wt.% GNPs at fracture
surface of Mode-I fracture toughness testing A) near the inserted film and B) far
from the inserted film. The red lines represent the agglomeration regions of GNPs

234
The dominant feature of the failure of TPU70/CF laminates and MSCs is different.
While the TPU70/CF and MSCs combined with 0.5 wt.% GNPs show a river pattern
and a large amount of hackle pattern, a very small amount of hackle and/or river
pattern can be observed in the MSCs incorporated with 2.5 and 5 wt.% GNPs,
respectively as illustrated in Figure 6.7, Figure 6.8, Figure 6.9 and Figure 6.10
respectively. The river pattern is attributed to the failure mode when the matrix
surrounds the fibre and indicates the plastic flow of the matrix [360-363], whereas
the hackle pattern is a characteristic of shear deformation [364].

Unfortunately, very limited work has been performed on the interlaminar fracture
toughness of MSCs based on thermoplastic/CF composites. Gabr et al. [304] studied
the effect of the addition of nanoclay on the ILFT of polypropylene containing
compatibilizer (PPC) and reinforced by CF . Their study revealed that incorporating
3 wt.% organoclay within PPC using a melt-compound process caused an
enhancement in the initiation and propagation of GIC by 64% and 67%, respectively.
Their results are incompatible with the present results; this may be because of using a
compatibilizer with PP in order to increase interfacial adhesion between PP and
nanoclay from one side and PP/NCs and CF from another side. The SEM micrograph
of their study observed a rough surface of PPC/CF incorporated with nanoclay,
suggesting a high delamination resistance. In spite of the differences in the
explanation of the SEM micrograph observations between the current study and that
reported by Mohamed et al. [304], the observation of increase fracture surface
roughness due to the incorporation of nanoparticles is the same.

In order to see if the reduction in ILFT of TPU70/CF due to the addition of high
loading GNPs is still within the range and does not deteriorate the ILFT of the
resultant MSCs, it is necessary to establish a comparison between the present results
and those of other studies of epoxy/CF modified with different nanoparticles. Table
6.3 summarises some results of GIC-Max and GIC-Prop of epoxy/CF multiscale
composites for comparison with the current results, which are tabulated in Table 6.2.
It can be clearly seen that the reduction in ILFT of MSCs incorporated with 5 wt. %
GNPs is still within the range of ILFT of epoxy/CF multiscale composites. This
means that the composites or MSCs based on TPU-70 HS and consolidated with
GNPs produce an ILFT that is greater than or comparable with that of composites
based on epoxy matrices.

235
Table 6.3: Average values of GIC-Max and GIC-Prop of different epoxy/CF multiscale
composites

GIC-Max GIC-Prop
No. Authors Matrix Fibres Fillers Ref.
(kJ/m2) (kJ/m2)
1 1.5 wt%
Borowski
Epoxy CF (COOH- 1.12 - [127]
et al.
MWCNTs)
2 1 wt.%
Wang et CF- carbon
Epoxy - 0.285 [137]
al. fabric nanofibre
(CNF)
3 Yokozeki Cup-stacked
Epoxy CF 0.25 [138]
et al. CNTs
4 Ashrafi et 0.1 wt.%
Epoxy CF 0.314 0.343 [365]
al. SWCNTs
5 TPU-70 5 wt. % Present
Naji 1.60.1 1.20.1
HS GNPM25 study

6.3.2 Impact Damage Resistance

To study the effect of adding GNPs to a TPU-70 HS matrix on the impact


performance and damage tolerance of TPU70/CF laminates and MSCs, low-velocity
impact tests were carried out. In order to conduct this test, a drop-weight impact-
testing machine was employed and was calibrated to impact energy levels of 5, 15
and 25 J. The response of each sample was recorded in terms of load, deflection and
energy. The typical curves of force versus time, force versus displacement and
energy versus time at different impact energy levels are presented in Figure 6.11,
Figure 6.12 and Figure 6.13, respectively. A comparison of the results of peak force,
peak deformation and absorbed energy for TPU70/CF and MSCs at applied impact
energy levels are shown respectively in Figure 6.11 D, Figure 6.12 D and Figure 6.13
D. The corresponding data of peak load, peak deformation and absorbed energy of
TPU70/CF laminates and MSCs are tabulated in Table 6.4.

Generally, the contact force is defined as the compressive forces applied by the
specimens to the impactors [366]. Figure 6.11 and Figure 6.12 illustrate the contact
force-time and contact force-displacement curves of TPU70/CF and MSCs at various
energy levels respectively. From these figures, the first sudden drops and/or the
variation in the slope of the ascending linear portion of force-time and force-
displacement curves are denoted as incipient or initial damage load (Pi). This Pi

236
typically results from internal delamination and/or failure in the fibre-matrix
interface, which is normally located near the rear surface of the impacted specimen
and forms during the first stage of impact events [367-369]. Meanwhile, the
maximum point of contact force (the force between the impactor and the sample)
represents the maximum load (Pm) that the composites can tolerate before major
damage occurs. This major damage normally begins at the rear surface (tension area)
due to the fibre breakage, which eventually leads to the occurrence of penetration on
the front surface (compression area). Hence, the damage can continue in different
failure modes, such as tension, compression, or even the transverse shear mode,
which deteriorates the stiffness of the composites [366-369].

(A) (B)
5J Maximum Force
15J
5000 8000 TPU70/CF
TPU70/CF (Pm)
TPU70/0.5wt.%GNP/CF
TPU70/0.5wt.%GNP/CF TPU70/2.5wt.%GNP/CF
4000 TPU70/2.5wt.%GNP/CF TPU70/5wt.%GNP/CF
TPU70/5wt.%GNP/CF 6000
Force (N)

Force (N)

3000

4000
2000

Pi Pi
2000
1000

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6
Time (ms) Time (ms)
(C) (D)
25J
10000 TPU70/CF 10000
TPU70/CF
TPU70/0.5wt.%GNP/CF
TPU70/0.5 wt.%GNP/CF
TPU70/2.5wt.%GNP/CF
8000 TPU70/2.5 wt.%GNP/CF
TPU70/5wt.%GNP/CF
TPU70/5 wt.%GNP/CF
Peak Force (N)

8000
Force (N)

6000

6000
4000

4000
2000

0 2000
0 1 2 3 4 5 6 7 5 10 15 20 25
Time (ms) Impact Energy (J)

Figure 6.11: The response of contact force versus time of TPU70/CF laminate and
multiscale composite incorporated with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at
various impact energy levels: A) 5 J, B) 15 J and C) 25 J, and D) the comparison of
peak force at these impact energy levels
As can be clearly seen from Figure 6.11, the contact force of TPU70/CF laminates
and MSCs increase as the impact energy levels increase, recording its highest value

237
at 25J impact energy. Regardless of the type of samples, the trends of the force-time
and force-displacement curves of MSCs are similar to those of the TPU70/CF
laminates. It can also be observed from Figure 6.11 that the force-time curves of
TPU70/CF laminates show a slightly higher peak than the other MSC specimens,
which is designated as contact stiffness [145, 370, 371]. In fact, the TPU70/CF and
MSCs show a clear variation in the initial slope, which is denoted as Pi, before
gradually rising to reach a maximum value of contact force (Pm). The addition of
GNPs to the TPU70/CF causes a slight reduction in Pi value from 541.31  53 N of
TPU70/CF to 501  36 N, 348  33N, and 538  75 N of MSCs combine with 0.5,
2.5 and 5 wt.% of GNPs, respectively. This result might be assigned to the reduction
in strength retention of TPU70/CF due to its incorporation with GNPs with high
stiffness [10].

To be precise, after Pi, the contact force curves of TPU70/CF and the MSCs
experience various modes of short oscillation or a sawtooth plateau that can be
attributed to the formation of matrix damage at different impact energy levels [144,
145]. The value of these oscillations is comparable to the size and/or severity of the
damage [144]. The same behaviour is observed in force-displacement curves, as
illustrated in Figure 6.12. After short oscillation regions, the force-time response of
all sample sets exhibit neither a drop nor slope variations until reaching Pm,
indicating the elastic response up to this point [370, 372]. The sets of samples that
did not exhibit this behaviour are the TPU70/CF laminates at 25 J, which undergo
some oscillation in a load before reaching Pm (see Figure 6.11 C). This result may
indicate the penetration of an impactor on the impacted surface of the specimen [370,
372]. This phenomenon was not observed in the MSC samples, as illustrated in
Figure 6.11 C. After attaining Pm, the contact force decreases smoothly until it
becomes equal to zero or approaches the minimal value. This trend of load reduction
or the symmetry of ascending and descending parts of load-time curves indicates
qualitatively that the creation of damage is not sufficiently extensive to cause a sharp
reduction in the stiffness of the laminates [144, 370].

In general, the addition of GNPs to the TPU70/CF reduces their toughened system;
this claim can be confirmed by the slight reduction in Pm shown in Figure 6.11 A, B,
C, and D and Table 6.4. For example, at 25 J impact energy, Pm drops from 9571 
85 N of TPU70/CF laminates to 8096  238 N, 7980  186 N, and 8855  10 N for

238
MSCs combined with 0.5 wt.%, 2.5 wt.% and 5 wt.% of GNPs respectively.
However, after reaching Pm, the contact force of TPU70/CF and MSCs drops
gradually at all impact energy levels.

As stated previously, there is no sudden decrease in load after Pm which can be


attributed to the severe damage due to the brittleness of the matrix. Despite the
addition of GNPs, all of the TPU70/CF laminates and MSCs failed in the ductile
mode. This finding can be interpreted in two ways: firstly, it may be that the creation
of damage is not enough to cause a significant reduction in composite stiffness.
Secondly, the low weight percentage of GNPs does not cause a significant reduction
in the toughness and ductility of TPU70/CF laminate. As demonstrated earlier in this
study in Chapter four (see sections 4.3.4.3, 4.3.6.1 and 4.3.6.2) , the presence of
GNPM25 in TPU-70 HS leads to an acceleration of the formation of micro-
crystalline HD and a subsequent improvement in the microphase separation of SP
and HP, resulting in an improvement in the stiffness of NCs. Moreover, it was also
proven earlier that incorporating GNPM25 with TPU70/CF to form MSCs reduced
the interfacial interaction between the NC matrix and CF (see chapter 5). This
interpretation could explain the reduction in the initial slope of force-time and force-
displacement curves, as illustrated in Figure 6.11 and Figure 6.12.

Like the force-time curves, the force-displacement curves in Figure 6.12 undergo a
short oscillation trend after passing Pi, which changes from one impact energy level
to another for the TPU70/CF laminates. It can be unambiguously observed from this
figure that at the loading phase of the curve for the TPU70/CF laminates at 5 J
impact energy the force increases linearly with displacement until it reaches the
maximum value, while at 15 and 25 J the curves show a subtle continuous oscillation
trend or sawtooth profile. The sawtooth profile is attributed to the impact damage
caused to the matrix. On the contrary, the contact force of MSCs, regardless of their
loading content, increases linearly with displacement until reaching Pm at all impact
energy levels. This conclusion indicates that the presence of GNPs in TPU70/CF
reduces the damage and fibre breakage by improving the damage resistance of the
matrix. Once Pm is reached, the force-displacement curves tend to behave in a
nonlinear way in the unloading region.

239
(A) (B)
TPU70/CF 5J TPU70/CF 15J
5000 TPU70/0.5wt.%GNP/CF
Pm 8000 Peak Deformation
TPU70/0.5wt.%GNP/CF
TPU70/2.5wt.%GNP/CF TPU70/2.5wt.%GNP/CF
TPU70/5wt.%GNP/CF TPU70/5wt.%GNP/CF
4000
Force (N) 6000

Force (N)
3000

4000
2000
Pi Pi
2000
1000
Permenant Deformation

0 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 0 1 2 3 4 5
Displacement (mm) Displacement (mm)
(C) 7
(D)
TPU70/CF
10000 TPU70/0.5wt.%GNP/CF
25J
TPU70/2.5wt.%GNP/CF
6 TPU70/CF
TPU70/5wt.%GNP/CF
TPU70/0.5 wt.%GNP/CF

Peak Deformation (mm)


8000 TPU70/2.5 wt.%GNP/CF
TPU70/5 wt.%GNP/CF
5
Force (N)

6000
4
4000
3
Pi
2000
2
0
0 1 2 3 4 5 6 7 5 10 15 20 25
Displacement (mm) Impact Energy (J)

Figure 6.12: The corresponding contact force versus displacement of TPU70/CF


laminates and multiscale composites combined with 0.5 wt.%, 2.5 wt.% and 5 wt.%
GNPs at different impact energy levels: A) 5 J, B) 15 J and C) 25 J, and D) the
comparison of peak deformation at these impact energy levels
An analogous plateau is noticed for all the MSC samples, regardless of their GNP
loading content. The same load-displacement behaviour in epoxy/CF laminates
modified with hybrid nanofillers consisting of MWCNT and boron nitride was
observed by Hasan Ulus et al. [145], despite the difference with this work in matrices
materials. This research was mentioned previously in a brief way at the literature
review chapter (chapter 2, section 2.4.4.4).

On the other hand, the peak deformation values also increase with increasing impact
energy levels, as seen in Table 6.4 and Figure 6.12. For example, TPU70/CF
laminates exhibit a maximum central deflection of around 2.17 mm at an impact
energy of 5 J, while it reaches 4.48 mm at 25 J impact energy. The smallest disparity
between TPU70/CF and MSCs is recorded at 5 J impact energy. This result means
that the 5 J impact energy is not adequate to demonstrate the differences between
TPU70/CF laminates and MSCs in stiffness and deformation [47].

240
These types of curves can identify the types of responses of laminates that occur
during subjection to impact loading, such as penetration, perforation and rebounding
[366, 373, 374]. It can also be seen from Figure 6.12 and Table 6.4 that the
TPU70/CF shows a slightly lower peak deformation than the MSCs. In spite of the
slight increase in peak deformation and a modest reduction in Pm, as shown in Table
6.4 and as a consequence of the addition of GNPs, the curve is still a closed type,
which indicates that no perforation has taken place within the composite specimens.
In general, the initial slope of force-displacement curves is denoted as impact-
bending stiffness of laminated composites throughout the impact events [371, 374].
Therefore, a slight reduction in the initial slope might refer to the reduction in
bending stiffness of the composites during impact events owing to the weak bonding
between the CF and NCs matrices, as mentioned previously.

The energy-time response curves of TPU70/CF laminates and MSCs at different


energy levels such as 5, 15 and 25 J are displayed in Figure 6.13. In this figure, the
peak point represents the total impact energy (Et). The impact energy gradually
decreases over time once the maximum energy point is reached; this reduction is
attributed to the recovery of elastic energy or rebound energy. The lowest value in
these curves represents the absorbed energy, which denotes the amount of energy
absorbed by the specimen at the end of the impact events [366]. The rebound energy
(Er) or the amount of elastic energy that returns to the impactor to cause rebounding
can be calculated from the difference between the highest point (Et) and the lowest
point (Ea). This means that at perforation state Er will be equal to zero. The absorbed
energy was calculated automatically by the software according to the following
equation:
𝟏
𝑬𝒂 = 𝟐 𝒎(𝒗𝟐𝒊 − 𝒗𝟐𝒓 ) ....... ( 6.11)

where:

Ea - absorbed energy in J

M - mass of the impactor in kg

vi - velocity of the impactor in m/s

vr - velocity of the rebounding impactor in m/s

241
(A) (B)
6 16
5J
14 15J
5
Rebonded
12 Energy(Er)

Energy (J)
4
Energy (J) 10

3 8

6
2
TPU70/CF A bsorbed energy (Ea) due to damage
4
TPU70/0.5wt.%GNP/CF
1
TPU70/2.5wt.%GNP/CF 2
TPU70/5wt.%GNP/CF
0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Time (ms) Time (ms)
(C) (D)
25 25J 16 TPU70/CF
TPU70/0.5 wt. %GNP/CF
14 TPU70/2.5 wt. %GNP/CF

Absorbed Energy (J)


20 TPU70/5 wt. %GNP/CF
12
Energy (J)

15 10

8
10
6
TPU70/CF
TPU70/0.5wt.%GNP/CF 4
5 TPU70/2.5wt.%GNP/CF
TPU70/5wt.%GNP/CF 2

0 0
0 1 2 3 4 5 6 7 5 10 15 20 25
Time (ms) Energy Level (J)

Figure 6.13: The response of energy versus time of TPU70/CF laminate and
multiscale composite combined with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at various
impact energy levels: A) 5 J, B) 15 J and C) 25 J, and D) the comparison of
absorbed energy at these impact energy levels
As can be clearly seen from Figure 6.13 A, B, C and D and Table 6.4, the absorbed
energy of the TPU70/CF laminates and MSCs increases with increasing impact
energy levels; this will result in an increase in the severity of the damage to the
specimens [375], while the rebound energy decreases. It can also be seen from Table
6.4 that the Ea for all four samples is lower than the total impact energy, indicating
that no penetration or perforation has occurred. For instance, the TPU70/CF samples
absorbed 2 J for impact energy of 5 J and 13.5 J for an impact energy of 25 J. The
MSCs absorbed high energy at all impact energy levels, especially at 15 J and 25 J,
since at 5J impact energy there is a marginal difference between the absorbed energy
of the TPU70/CF and MSCs. Meanwhile, at 15 J and 25 J, the MSC samples,
especially MSCs combined with 2.5 and 5 wt. % GNPs, absorbed high values of
energy compared with TPU70/CF. This result implies that the MSCs have the ability
to absorb more energy at high energy levels (15 and 25 J) than the TPU70/CF, due to
their combination with GNPs. Additionally, the high amount of energy absorbed by

242
the MSCs is due to the extensive energy that is consumed to generate various
damage failures such as delamination, fibre breakage and matrix-fibre debonding.
Therefore, the MSC samples would also have a high dent depth and high damage
area, as was mentioned previously and will be confirmed later in this thesis. On the
other hand, the low absorbed energy of TPU70/CF compared with MSCs, at all
energy levels, can indicate good adhesion between the TPU-70 HS and CF. This
means that the composites with good adhesion can sustain a higher load to create
damage and lower absorbed energy, especially at 5 J impact levels.

Table 6.4: Results of low-velocity impact test of TPU70/CF laminates and MSCs at
three different impact energy levels (5 J, 15 J and 25 J)

Impact Peak Rebound Absorbed


Peak Force
Sample Energy Deformation Energy Energy
(N)
(J) (mm) (J) (J)
TPU70/CF 4489  245 2.2  0.2 2.9  0.2 2.0  0.2
TPU70/0.5
3935  246 2.4  0.2 2.9  0.2 2.1 0.2
wt.%GNP/CF
TPU70/2.5
5J 3346  53 2.8  0.1 2.4 0.1 2.6  0.1
wt.%GNP/CF
TPU70/5
3525  153 2.5  0.2 2.4  0.2 2.6  0.2
wt.% GNP/CF
TPU70/CF 7425  584 3.7  0.4 7.6  0.5 7.3  0.5
TPU70/0.5
6953  260 4.1  0.2 6.9  0.3 8.0  0.3
wt.%GNP/CF
TPU70/2.5
15 J 6193  39 4.7  0.1 5.9  0.3 9.0  0.3
wt.%GNP/CF
TPU70/5
6461  205 4.3  0.2 6.3  0.3 8.6  0.3
wt.%GNP/CF
TPU70/CF 9571  85 4.5  0.1 11.3  0.1 13.5  0.1
TPU70/0.5
8096  237 5.9  0.2 9.3  0.3 15.6  0.3
wt.%GNP/CF
TPU70/2.5
25 J 7980  186 6.3  0.2 8.9  0.2 15.9  0.1
wt.%GNP/CF
TPU70/5
8855  10 5.5  0.1 9.5  0.4 15.4  0.4
wt.%GNP/CF

In spite of the differences in the nature of the matrices, similar behaviour was
observed by Mahdi et al. [142] of epoxy/CF laminates incorporating 2 wt.%
nanoclay, 0.3 wt.% MWCNT and a hybrid of 0.1 wt.% MWCNT with 2 wt.%
nanoclay. They found that the epoxy/CF samples yielded lower absorbed energy
compared to samples reinforced with nanoclay and hybrid nanoparticles. They
revealed that the samples of composites modified with hybrid nanoparticles exhibited

243
greater absorbed energy, peak load and lower deflection compared with other
samples. They believed that the reason for these results could be attributed to the
ability of MWCNTs and nanoclay to form a synergistic effect, since they both help to
transfer loads between the matrix and fibres. Their study is similar to another
previous study [143].

From the literature review, Eskizeybek et al. [144] revealed that the addition of
CaCO3 nanoparticles to epoxy/CF caused a decrease in Pm and an increase in
absorbed energy and peak deflection at 3 wt.% loading, while 2 wt.% and 1 wt.%
caused unexpected growth in peak load and drops in absorbed energy. They
attributed this result to the ability of nanoparticles to obstruct crack growth when the
crack comes into contact with them, meaning that the composite samples needed to
absorb more energy to create and/or propagate the damage [140].

By using a C-scan, the damaged areas of the specimens after being subjected to an
impact test can be identified and examined so as to establish a relationship between
the impact energy and impact resistance. As mentioned previously, an increase in
the absorbed energy by the addition of GNPs causes an increase in the damage area,
as was confirmed by an ultrasonic C-scan test. Therefore, the relationship between a
damage area measured using ImageJ software and the impact energy levels are
illustrated in Figure 6.14.

1200
TPU70/CF
1000 TPU70/0.5 wt.% GNP/CF
TPU70/2.5 wt.% GNP/CF
Damage Area (mm2)

TPU70/5 wt.% GNP/CF


800

600

400

200

5J 15 J 25 J
Impact Energy (J)

Figure 6.14: Damage area measurements of TPU70/CF and MSCs combined with
0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at various impact energy levels

244
Three observations can be made from this figure; firstly, the damage area for all
specimens grows as the impact energy levels increase. Secondly, the TPU70/CF
laminates exhibit a smaller damage area than the MSCs at all impact energy levels.
Thirdly, the damage area of MSCs increases as the weight percentage of GNPs
increases at the same impact energy levels. For example, at an impact energy of 5 J
the damage area of MSCs combined with 0.5 wt.% GNPs is 55 mm2, which increases
to become 150 mm2 and 317 mm2 of MSCs with 2.5 wt.% and 5wt.% GNPs
respectively. This indicates that the presence of GNP flakes provides less resistance
to impact damage and ensuing propagation of cracks due to its brittle nature. In
addition, this finding might be attributed to the agglomeration or poor distribution of
GNPs, at high loading, which act as flaws. These flaws act as high stress-
concentration points. As a result, facilitating crack progression takes place and
additional damage occurs. These results were confirmed by Mode-I ILFT results,
where the crack initiation and propagation was fast and catastrophic for samples
containing higher levels of GNP loading due to agglomeration, which is considered a
major cause of reduction.

A further explanation may be that the energy imparted by the impactor to the
impacted sample surface was dissipated around the contact area of the MSC samples;
due to the poor interfacial adhesion with CF, severe delamination and splitting will
occur. Meanwhile, the good adhesion of TPU70/CF suggests a concentrated energy
point, leading to the occurrence of local damage [10]. The reason suggested
previously by Eskizeybek et al. [144] may also apply here. However, an analogous
explanation of nanoparticle agglomeration was suggested by Hosur et al. [370], who
modified epoxy/CF with nanoclay. They found that the addition of 3 wt.% nanoclay
to epoxy/CF composite resulted in a significant increase in the damage area
compared with 1 wt.% and 2 wt.%. They attributed this high growth in the damage
area to the occurrence of nanoclay agglomeration at high loading.

The dent in the impacted surface of the composite samples is caused by the impact
test. The depth of the dent can be measured using a digital gauge directly after the
test; in this situation, the result represents the temporary dent depth. Measurement is
then repeated after 24 hours to give the permanent dent depth. In fact, the dent depth
measurements can refer to the response of the composites adjacent to the point
between the impactor and the contact surface of the composite samples. The

245
measurements of the dent depth of all four samples at various impact energy levels
are compared and illustrated in Figure 6.15.
(A) (B)
0.40
0.18 5J Temporary 15 J
Temporary
0.35 Permanent
0.16 Permanent

0.14 0.30
Dent depth (mm)

0.12 0.25

0.10 0.20
0.08
0.15
0.06
0.10
0.04
0.05
0.02

0.00 0.00
/CF /CF F /CF /CF F
GNP GNP NP/C F GNP GNP NP/C
U70
/CF wt.% wt.% t.%G U 70/C wt.% wt.% wt.%G
TP 0/0.5 0/2.5 70/5 w TP
70/0
.5
70/2
.5 70/5
TPU
7
TPU
7 TPU TPU TPU TPU

(C)
Temporary 25 J
0.5 Permanent

0.4
Dent depth (mm)

0.3

0.2

0.1

0.0
/CF /CF F
F GNP GNP NP/C
P U 70/C .5 wt.% .5 wt.% wt.%G
T 70/0 70/2 70/5
TPU TPU TPU

Figure 6.15: Comparison of dent depth recovery measurements of TPU70/CF


laminates and MSCs incorporated with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs at
impact energy levels of A) 5 J, B) 15 J and C) 25 J respectively
As with the damage area curves, three important observations can be made from this
graph. Firstly, the dent depth increases with increasing impact energy of all four
samples and the greatest dent depth is evidenced in MSCs combined with 5 wt. %
GNPs. Secondly, the TPU70/CF and MSCs combined with 0.5 wt. % GNPs have a
higher recovery at impact energy levels of 15 J and 25 J than that of the MSCs
combined with 2.5 wt.% and 5 wt.% GNPs. This result can be attributed to the ability
of the TPU70/CF and MSCs combined with 0.5 wt.% GNPs to substantially recover
the local deformation that was introduced at high impact energy, while at low impact
a very small dent formed on the surface of both samples. Meanwhile, the MSCs
combined with 2.5 wt.% and 5 wt.% GNPs exhibit low dent-depth recovery,
suggesting localised sub-surface modes such as delamination [376]. It can also be
observed from this graph that the dent-depth recovery decreases with increasing GNP

246
content at high impact energy levels. This implies that the presence of a high loading
of GNPs helps to form a large amount of delamination that remains during impact
events.

6.3.2.1 Impact Damage Evaluation using SEM


The SEM cross-section images of TPU70/CF laminates and MSCs after subjection to
impact energy of 25 J are shown in Figure 6.16. During an impact event, the
composite specimen absorbs the impact energy via elastic and plastic deformation
and through various damage modes. These damage modes represent impact damage
mechanisms, which consist of delamination, debonding and fibre breakage [6, 8,
370]. In general, for all of the images shown in Figure 6.16, the damage occurs on
the back surface of the samples which is far from the impactor surface. The load was
transferred from the impacted surface of the specimens to the other side. The
impacted specimen of TPU70/CF laminates in Figure 6.16 A shows a low-impact
damage mode. The samples responded to the impact energy in the form of slight
delamination, fibre-matrix debonding, and a tension failure at the bottom layer. This
response explains their low damage area and dent depth.

Conversely, as seen in Figure 6.16 B, MSCs combined with 0.5 wt.% GNPs show the
same slight impact damage as the TPU70/CF laminates, as well as matrix cracks.
These types of damage are not as severe as those found in the MSCs combined with
2.5 wt.% and 5 wt.% GNPs. It can be seen from Figure 6.16 C and D that the MSCs
combined with 2.5 wt.% and 5 wt.% GNPs exhibit extensive impact damage, as seen
in the significant amount of delamination, matrix cracks, fibre-matrix debonding,
fibre breakage and in the large amount of debris located on the tension surface in the
impact line. In addition, both the MSCs combined with 2.5 wt.% and 5 wt.% GNPs
show a hole owing to the fibre being pulled out in the regions close to the bottom
layer, while this type of damage was not observed in the TPU70/CF and MSCs
combined with 0.5 wt.% GNPs samples. The reasons for these types of failure mode
can be attributed to the spread impact load between the NCs and the CF layer, in
spite of the weak fibre-matrix adhesion in the MSCs with large amounts of GNPs
that stimulate fibre-matrix debonding. It is also possible that the fibre breakage and
pull-out may be the reason for the high value of absorbed energy of this type of
MSCs, as suggested previously.

247
The same observation was reported by Junid [10]. He suggested that the addition of
A-GNPs to epoxy/CF helped to dissipate the energy imparted by impactors to the
surrounding area of the contact point between the impactor and the specimen, leading
to severe damage such as splitting or delamination due to the poor interfacial
bonding between NC matrices and CF. On the contrary, the epoxy/CF sample
presents concentrated energy at the point of contact, leading to the creation of an
intralaminar crack and a small damage area.

A Impact direction

Delamination

Fibre-Matrix 100 m
Matrix crack Debonding

B
Matrix
crack
Delamination

100 m

C
Delamination Fibre-Matrix Debonding

100 m
Debris

248
D

100 m

Figure 6.16: SEM cross-section images after subjection to 25 J impact energy of A)


TPU70/CF laminates and B) MSCs incorporated with 0.5 wt.% GNPM25 , C) MSCs
incorporated with 2.5 wt.% GNPM25 and D) MSCs incorporated with 5 wt.%
GNPM25

6.3.3 Compression after Impact (CAI)

Impacted and non-impacted TPU70/CF and MSC samples were subjected to


compression stress tests so as to study the influence of the addition of GNPs on their
damage tolerance. It is well known that impacted laminates exhibit a considerable
reduction in strength during loading by compression, owing to the local instabilities
arising from the existence of damage [141]. The damage tolerance can be determined
by evaluating the residual compressive strength of both the reference TPU70/CF
composites and MSCs. The residual compressive strength, on the other hand,
represents the ratio of the ultimate compressive strength of the composite samples
with impact damage to its value without impact damage.

The compression stress-extension curves of all four samples at each impact energy
level (0 J, 5 J, 15 J and 25 J) are plotted in Figure 6.17. A comparison of the ultimate
compression and CAI strength and residual compressive strength of all four samples
at different impact energies are illustrated in Figure 6.18 A and B respectively. The
corresponding data of the compression and CAI strength and residual compressive
strength (%) of all four samples with varying impact energies are tabulated in Table
6.5. It can be clearly seen from Figure 6.17 A, B, C and D that the compressive
strength of TPU70/CF and MSCs decreases with increasing impact energies as a
consequence of increasing the damage area.

249
120
(A) 120
(B)
0J 5J
100 100

Compression stress (MPa)


TPU70/CF
TPU70/0.5wt.%GNP/CF
80 80 TPU70/2.5wt.%GNP/CF
TPU70/5wt.%GNP/CF

60 60

TPU70/CF
40 40
TPU70/0.5wt.%GNP/CF
TPU70/2.5wt.%GNP/CF
20 TPU70/5wt.%GNP/CF 20

0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

100
(C) (D)
15 J 25 J
TPU70/CF 80
Compression stress (MPa)

80 TPU70/0.5wt.%GNP/CF TPU70/CF
TPU70/2.5wt.%GNP/CF TPU70/0.5wt.%GNP/CF
TPU70/5wt.%GNP/CF 60 TPU70/2.5wt.%GNP/CF
60 TPU70/5wt.%GNP/CF

40
40

20 20

0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Extension (mm) Extension (mm)

Figure 6.17: Compression stress versus extension curves of TPU70/CF laminates


and MSCs combined with 0.5 wt. %, 2.5 wt. % and 5 wt. % GNPs after various
impact energy levels: A) 0 J, B) 5 J, C) 15 J and D) 25 J, respectively
For conventional TPU70/CF samples, there is a slight reduction in CAI strength
compared to post-impact compressive strength at 5 J and 15 J impact energies, while
at 25 J the reduction in compressive stress rises. The amount of this reduction is
approximately 9%, 13% and 35%, observed at impact energy levels of 5 J, 15 J and
25 J respectively, as illustrated in Figure 6.18 and Table 6.5. These results imply that
the increase in impact energy causes an increase in delamination and matrix
cracking, resulting in a reduction in the CAI strength [376] (see Figure 6.16 A). In
addition, for MSCs, the reduction in compressive strength when compared with non-
impacted samples increases as the impact energy increases. This is a common finding
that has been attributed to the high stress concentration generated around the damage
region at high energy levels, leading to high local instability during the compression
test [10]. Meanwhile, the failure behaviour of the TPU70/CF curves experience a
sudden drop after attaining ultimate compressive strength for all impact energies, i.e.
0, 5, 15 and 25 J. On the contrary, the MSCs exhibit a gradual drop after the
maximum point, which extends with the increase in GNP concentration. The
extension in the drop region indicates an increase in plastic deformation behaviour,

250
suggesting that microcrack slippage might ensue during the compression test; a
similar suggestion has been proposed in another study [10].

Table 6.5: Compressive strength and residual compressive strength of TPU70/CF


and MSCs before and after impact testing at 0 J, 5 J, 15 J and 25 J.
Impact Sample Compressive Residual Compressive
Energy Strength (MPa) Strength (%)
(J)
TPU70/CF 112.2  3.0 100.0
TPU70/0.5 wt.%GNP/CF 88.0  2.5 100.0
0J TPU70/2.5 wt.%GNP/CF 69.3  5.2 100.0
TPU70/5 wt.%GNP/CF 70.0  3.5 100.0
TPU70/CF 101.7  20.8 90.6
TPU70/0.5 wt.%GNP/CF 84.5  0.5 96.0
5J TPU70/2.5 wt.%GNP/CF 56.0  2.0 80.8
TPU70/5 wt.%GNP/CF 56.0  7.0 88.0
TPU70/CF 93.6  1.6 83.5
TPU70/0.5 wt.%GNP/CF 73.5  1.5 83.6

15 J TPU70/2.5 wt.%GNP/CF 51.5  0.5 73.6


TPU70/5 wt.%GNP/CF 40.5  2.5 58.4
TPU70/CF 72.5  8.5 64.6
TPU70/0.5 wt.%GNP/CF 51.5  5.5 58.5
25 J TPU70/2.5 wt.%GNP/CF 40.0  0.8 57.1
TPU70/5 wt.%GNP/CF 30.5  3.5 44.1

(A)
160
TPU70/CF
140 TPU70/0.5wt.%GNP/CF
Compressive Strength (MPa)

TPU70/2.5wt.%GNP/CF
120 TPU70/5wt.%GNP/CF

100

80

60

40

20

0
0 5 15 25

251
(B)

Residual compressive strength (%)


100

90

80

70

60
TPU70/CF
50 TPU70/0.5wt.%GNP/CF
TPU70/2.5wt.%GNP/CF
40 TPU70/5wt.%GNP/CF

30
0 5 15 25
Impact Energy (J)

Figure 6.18: A) Ultimate compressive and CAI strength and B) residual compressive
strength (%) of TPU70/CF and MSCs combined with 0.5 wt.%, 2.5 wt.% and 5 wt.%
GNPs with varying impact energies
The addition of GNPs also causes a significant reduction in the compressive and CAI
strength of conventional TPU70/CF, and this reduction grows as the GNP content
increases. For example, for MSCs that are incorporated with 0.5 wt. % GNPs, the
reduction in CAI strength is slight when compared with MSCs combined with 2.5
wt.% and 5 wt.% GNPs. This reduction is around 17%, 21% and 30% for impact
energies of 5, 15 and 25 J respectively in comparison with the CAI strength of
TPU70/CF samples at the same energy levels. This is due to the small area of impact
damage induced during the impact test, as illustrated previously (see Figure 6.14).
On the contrary, the CAI strength of MSCs that contain 5 wt.% GNPs decreases by
45%, 57% and 58%, compared with conventional TPU70/CF (see Figure 6.17 and
Table 6.5). This decrease is from 101 MPa of TPU70/CF to 56 MPa, 41 MPa and 31
MPa of MSCs incorporated with 5 wt. % GNPs at impact energies of 5, 15 and 25 J,
respectively. The progressive reduction in the CAI strength of MSCs combined with
2.5 wt.% GNPs and 5 wt.% GNPs respectively is a consequence of the increase in
the damage area related to delamination, fibre breakage and debris on the non-
impacted surface of the samples (see Figure 6.16 C and D).

252
For comparison, the residual compressive strength (in %) of conventional TPU70/CF
and MSCs at varying impact energies is plotted in Figure 6.18 B. It can be observed
from this figure that the residual compressive strength of TPU70/CF decreases with
increasing impact energy and GNP concentration. The only sample that is exempt
from this finding is the MSCs combined with 0.5 wt.% GNPs at 5 J, the residual
compressive strength of which is slightly higher than that of TPU70/CF. This is
because of its very small damage area and lower delamination compared with that
combined with 2.5 and 5 wt.% GNPs samples at the same impact energy. This might
also indicate that the lower amount of GNPs (0.5 wt.%) does not induce weak
bonding between NCs and CF compared to MSCs with 2.5 and 5 wt.% GNPs. In
spite of the differences between epoxy and TPU-70 HS, Junid [10] reported a similar
outcome for CF/epoxy modified with different concentrations of A-GNPs. He found
that the existence of 2 wt.% A-GNPs within/epoxy/CF laminates causes a 56.3%
reduction in their CAI strength.

Unfortunately, very few studies have evaluated the damage tolerance of MSCs.
Ashrafi et al. [365] demonstrated a 3.5% improvement in the CAI strength of
epoxy/CF composites modified with 0.1 wt.% SWCNTs. Yokozeki et al. [138] found
0% and 7% enhancements in the CAI strength of CF laminate modified by 5 and 10
wt.% of cup-stacked CNTs. Kostopoulos et al. [122] reported a 12-15% rise in the
CAI strength of MSCs combined with 0.5 wt.% MWCNTs for different impact
energies.

6.4 Summary

Three different percentages of GNPs were incorporated into TPU-70 HS using in-situ
polymerisation. The resultant nanocomposites were used to form MSCs reinforced
by 5-harness-satin CF. The fibre volume fraction of TPU70/CF and MSCs lay
between 43% and 46% with a small void content. The MSCs were investigated to
study the effect of the incorporation of GNPs on Mode-I interlaminar fracture
toughness (ILFT), impact-damage resistance and CAI strength of reference samples
of TPU70/CF. These three tests were chosen to characterise the effect of the
incorporation of GNPs in TPU70/CF laminates due to their importance in the
structural applications of composite materials. The most important conclusions
obtained from these tests can be summarised as follows.

253
➢ The Halpin-Tsai and MH-T in micromechanical modelling proved that the
elastic modulus of the experimental data of TPU70-NCs combined with a
weight percentage of  1 wt.% of GNPM25 fits better with the Halpin-Tsai
model than with the MH-T in 2D or 3D directions.

➢ The Mode-I ILFT test revealed that the addition of a small amount of GNPs
(0.5 wt.%) caused a moderate increase in GIC-NL and GIC-VIS and a slight
enhancement in GIC-Max and GIC-Prop compared to TPU70/CF. This
improvement measured around 23% and 44%, of GIC-NL and GIC-VIS
respectively, and 10% of GIC-Max and GIC-Prop in comparison with TPU70/CF
respectively. Meanwhile, a high loading of GNPs caused a significant
decrease in ILFT, especially of GIC-Max and GIC-Prop. This is because of
agglomeration, which can act as a stress-concentration point and facilitate
crack initiation and growth, as demonstrated by SEM micrograph images.

➢ The testing of impact-damage resistance found that the incorporation of 0.5


wt.% GNPs did not cause a high increase in absorbed energy, dent depth or
damage area in comparison with TPU70/CF and other MSCs. In fact, the
absorbed energy of all three types of samples is higher than that of
TPU70/CF, suggesting that the MSCs consume this energy to form extensive
damage failures such as delamination, fibre breakage and matrix-fibre
debonding. For example, at an impact level of 25 J, the enhancement in the Ea
of MSCs combined with 2.5 wt.% and 5 wt.% GNPs is 11% and 17%
compared to TPU70/CF. In addition, a slight increase in the dent depth and
damage area of MSCs combined with 0.5 wt.% GNPs was found in
comparison with TPU70/CF and other MSCs.

➢ Both TPU70/CF and TPU70/0.5 wt.%GNP/CF showed a superior recovery in


dent depth at 15 J and 25 J impact energy levels after 24 hours in comparison
with MSCs incorporated with 2.5 wt. % GNPs and 5wt. % GNPs. This
indicates their ability to recover from the local deformation created during a
high impact-energy test.

➢ CAI strength decreased for all four samples (TPU70/CF, MSCs combined
with 0.5 wt.%, 2.5 wt.% and 5 wt.% GNPs) with increasing impact energy
levels. This is because of the stress concentration created around the damage

254
area. The CAI strength exhibited a greater reduction in MSCs incorporated
with 2.5 wt.% GNPs and 5 wt. % GNPs compared to the values of the
undamaged samples and to the TPU70/CF laminate. For example, for MSCs
combined with 5 wt.% GNPs, the CAI strength decreased from 101 MPa of
TPU70/CF to 56 MPa, 41 MPa and 31 MPa at impact energies of 5, 15 and
25 J respectively. This is due to the increase in the damage area as a result of
the delamination, debris and fibre breakage at the non-impacted surface.
Therefore, the percentage of residual compressive strength decreases with an
increase in impact energy and GNP concentration, except in the case of
MSCs combined with 0.5 wt.% GNPs, which experience a slight increase at
low impact-energy levels in comparison with TPU70/CF laminates.

255
Chapter 7: Conclusions and Suggestions for Future
Work

7.1 Conclusions

7.1.1 Introduction

This study has focused on manufacturing a new multiscale composites (MSCs) based
on the thermoplastic polyurethane (TPU) with 70 wt.% hard segments, reinforced
with carbon fibre (CF) fabric and nanofiller (GNPs). To this end, the TPU-70 HS
was filled with 5 wt. % GNPs with different mean diameter sizes of 5, 15 and 25 m
and synthesised using in-situ polymerisation. This type of nanocomposite was
produced first in order to study the effect of GNP size on the thermal, electrical and
mechanical properties of neat TPU-70 HS. Both unfilled and filled TPU-70 HS were
then used to manufacture new MSCs by reinforcing the samples with 5-harness-satin
carbon fibre using a compression-moulding process. The mechanical, electrical and
thermal properties of MSCs based on different sizes of GNPs were investigated so as
to attain a better size of GNPs that matched with the CF and formed an innovative
MSC with superior properties and performance. A study of the effect of annealing
treatment on neat TPU-70 HS, TPU-70/NCs and MSCs was also performed. The best
MSCs were selected and different weight percentages of GNPs (0.5, 2.5 and 5) were
used in order to study the effect of the amount of GNPs on the Mode-I ILFT, impact-
damage resistance and damage tolerance of the TPU70/CF laminates. The following
conclusions are based on the results from the experimental work in this thesis.

7.1.2 Characterisation of Graphene Nanoplatelets

➢ SEM and TEM observations of GNPs (M5, M15 and M25) revealed that the
sonication process did not cause any change in the GNPs’ shape. This result
means that the GNP flakes preserved their folding and rolling up after sonication
due to the short time and low frequency that was adopted in this study. WAXD
demonstrated that the sonication process had no effect on the GNPs’
crystallographic plane; their large and small peaks were still located at the same
angles of 26.6 and 55°, respectively.

256
➢ The elemental analysis using CHNS and XPS techniques confirmed that the
percentages of atomic concentrations of O, C and N of the GNPs changed after
extraction from TPU-70 HS. More specifically, the changes indicated an increase
in the amounts of O and N, which might indicate interfacial interaction between
the GNPs’ surface functionality and isocyanate, urethane, hydroxyl or ether
groups in the TPU-70 HS structure via covalent and/or H-bonds.

7.1.3 Characterisation of TPU-70 HS Nanocomposites

➢ The cryogenic fracture surface of TPU-70 HS and NCs using the SEM technique
revealed a good dispersion of all GNP types in very similar images. The OM
technique confirmed the good dispersion observed using the SEM technique for
all GNP types; however, it highlighted a shining area in the images of GNPM15
and GNPM25, which can refer to the agglomeration regions due to their high
surface area compared to GNPM5. The WAXD test of the NC samples showed
that there was no shift in the GNPs’ peak, indicating that there is no change in the
GNPs’ crystallographic plane due to the extra exfoliation of the GNP layers
during NC synthesis.
➢ The study of the effect of GNP size on the thermal stability, melting and
crystallisation temperatures and dynamic properties of TPU-70 HS and TPU-
70/NCs was investigated using TGA, DSC and DMTA techniques. The TGA test
confirmed that the thermal stability of TPU-70 HS experienced greater
improvement upon incorporation with GNPM15 and GNPM25, due to their high
surface area and aspect ratio compared to GNPM5. Isothermal DSC results
showed that the addition of GNPs caused a slight increase in the Tm and TC of neat
TPU-70 HS and the greatest improvement was attributed to GNPM25. From the
DMTA results, the E’ of all NCs is higher than the TPU-70 HS; this is attributed
to the improvement in the stiffness of TPU-70 HS due to the high stiffness of
GNPs, which works as a reinforcing agent. In addition, the Tg of the SS and HS
shifts to a slightly higher temperature with the addition of GNPs, regardless of
their size. This result is attributed to the restriction of the molecular chain mobility
of TPU-70 HS and their frictional motion due to the interfacial interaction
between GNPs and the HS in the TPU-70 HS structure.
➢ The annealing treatment shows a slight reduction in thermal stability, Tm, TC and
dynamic properties of NCs compared to its value before annealing. The major
257
causes of this reduction can in all probability be attributed to the more significant
increase in filler-filler interaction than for filler-polymer interaction, leading to the
restacking/agglomeration of GNPs that prohibit the reordering of HS and then
decrease microphase separation. However, the annealing treatment showed a
moderate improvement in the thermal stability, TC, Tm and dynamic properties of
neat TPU-70 HS. This result indicates the increase in the phase separation of the
SS and HS and the reordering of HS, which causes this improvement.
➢ A melt rheology test was carried out to investigate the effect of GNP size on the
rheology properties of neat TPU-70 HS and its nanocomposites using the
oscillation amplitude and oscillation frequency. It was concluded that the
GNPM25 had a greater improvement in G’ and * due to its high surface area,
aspect ratio, and good dispersion compared with GNPM5 and GNPM15.
➢ The tensile and flexural tests of TPU-70 HS and NCs revealed an improvement
in the tensile and flexural modulus by the addition of GNPs, regardless of their
size. The greatest improvement was attributed to the NCs combined with
GNPM15 and GNPM25 in comparison with GNPM5. The annealing treatment
caused a slight reduction in the flexural and tensile modulus of NCs based on
GNPM15 and GNPM25, and a minimal increase of neat TPU-70 HS and NCs
based on GNPM5.
➢ The electrical properties of NCs based on GNPM15 and GNPM25 exhibited a
significant increase in electrical conductivity in the in-plane and out-of-plane
directions compared with neat TPU-70 HS. This result indicated that adding 5wt.
% GNPM15 and M25 reached its percolation threshold, while NCs incorporated
with the same weight percentage of GNPM5 still acted as insulators due to its
smaller size.
➢ The thermal conductivity of TPU-70 NCs incorporated with GNPM25 is high in
comparison with neat TPU-70 HS and NCs based on GNPM5 and GNPM15.
➢ Overall, the significant improvement in the electrical, thermal and mechanical
properties of TPU-70 HS is attributed to the NCs combined with GNPM15 or, in a
particular way, with GNPM25, due to their high surface area and aspect ratio,
which may help to increase the probability of interfacial interaction with the HS
of the TPU-70 HS structure.

258
7.1.4 Characterisation of Multiscale Composites

The conclusions from the reported results of the dynamic, mechanical, electrical and
thermal properties of TPU70/CF laminates and MSCs based on different sizes of
GNPs can be summarised as follows:

➢ The DMTA results for the MSCs revealed that the MSCs incorporated with
GNPM15 and GNPM25 had a greater improvement in E’ than GNPM5, in
comparison with TPU70/CF. The TgHP of the TPU70/CF laminates experienced a
considerable improvement with the addition of GNPs, regardless of their size.
However, the annealing treatment caused a moderate reduction in the E’ and TgHP
of TPU70/CF and MSCs due to the effect of GNPs on the formation of
microcrystalline HD and phase separation.
➢ Tensile and flexural tests showed a significant improvement in the flexural
modulus of TPU70/CF with the addition of GNPM15 and GNPM25 and a slight
improvement in the elastic modulus, E, of GNPM5. Conversely, annealing
treatment caused a reduction in the flexural and tensile properties of MSCs based
on GNPM25, but a slight increase in flexural and tensile properties of TPU70/CF
and MSCs based on GNPM5.
➢ The effect of the addition of GNPs on the interlaminar shear strength (ILSS) was
studied based on TPU70/CF and MSCs. In spite of the improvement in the
flexural properties of TPU70/CF before annealing due to the addition of GNPs,
the ILSS decreased unexpectedly for all MSCs, regardless of GNPs size. The
slight reduction was ascribed to MSCs based on the GNPM25/NCs, when ILSS
decreased by 12% compared to the same value for TPU70/CF.
➢ The electrical conductivity of TPU70/CF increased slightly in the in-plane
direction and significantly in the out-of-plane direction with the addition of
GNPM15 and GNPM25. The out-of-plane thermal conductivity of TPU70/CF was
improved by the addition of GNPs; unexpectedly, the greatest improvement was
ascribed to MSCs based on the GNPM5/NCs.
➢ From the results for NCs and MSCs based on the effect of different sizes of GNPs
on the thermal, mechanical and electrical properties, GNPM25 can be chosen as
the best candidate of GNP size to form MSCs. However, it exhibited a failure to
enhance the thermal conductivity and ILSS.

259
7.1.5 Damage Tolerance and Delamination Resistance of TPU70/CF
and MSCs

The MSCs based on GNPM25 were selected in order to study the effect of the weight
percentage of GNPM25 on the delamination resistance, impact- damage resistance
and damage tolerance of TPU70/CF. The selection was adopted according to the
ability of GNPM25 to bring about a greater improvement in the mechanical and
electrical properties of TPU70/CF laminates and the mechanical, electrical and
thermal properties of TPU-70 HS than GNPM5. The following conclusions were
reported according to the investigation of TPU70/CF laminates and MSCs using
Mode-I ILFT, low-velocity impact and CAI tests.

➢ The Mode-I ILFT showed a slight improvement in GIC-NL and GIC-VIS and a
moderate improvement in GIC-Max and average GIC-Prop of MSCs based on the 0.5
wt.%GNP/NC compared to traditional TPU70/CF. However, the higher loading of
GNPs caused a significant reduction of all three values of G IC initiation and GIC
propagation (GIC-Prop.). This is because of agglomeration, which acts as a stress-
concentration point and enables the fast initiation and growth of cracks, as
demonstrated by SEM micrograph images.
➢ Contrary to the improvement in Mode-I ILFT caused by the incorporation of 0.5
wt.% GNPs, the impact-damage resistance and CAI at high impact-energy levels
exhibited a slight reduction, while the CAI at 5 J showed a slight drop. Similar to
the effect of high loading of GNPs on the ILFT, both the impact-damage
resistance and residual compressive strength experienced a significant reduction
compared to TPU70/CF laminats. This can be attributed to the agglomeration of
GNP flakes at high loading, or may be due to weak interaction between NCs and
CF at a high loading of GNPs.

7.2 Suggestions for Future Work

Based on the conclusions of the experimental studies in this thesis, the following
suggestions for future work are recommended:

➢ During this study, the GNPs were used without any treatment; this decreased the
ability of the nanofillers to form a strong interfacial interaction with the TPU-

260
polymer matrix. Therefore, it is recommended to use treated GNPs to produce
TPU nanocomposites with optimised mechanical, dynamic and thermal properties.
➢ To reduce the cost and enhance the properties of TPU nanocomposites, hybrid
nanofillers are suggested. For example, using graphene oxide (GO) or reduced
GO with GNPs might help to improve the interfacial adhesion with matrices due
to improved filler-polymer interactions.
➢ The TPU nanocomposites based on the smallest size of GNPs (GNPM5) did not
show any enhancement in electrical conductivity; as such, using a high loading
of GNPM5 will help the nanocomposite to obtain an improvement in electrical
conductivity owing to reaching their percolation threshold.
➢ The MSCs based on GNPM5 showed a void content of approximately 8%,
which was higher than that of MSCs based on GNPM15 and GNPM25. Since
the high void content can reduce the properties of resultant composites, a
process of drying the nanocomposite at 80°C for one or two days instead of one
night is suggested, in order to attempt to reduce or avoid void content within
MSCs.
➢ The MSCs based on different sizes of GNPs showed a reduction in ILSS in
comparison with TPU70/CF laminates. Therefore, a pre-preg process is
suggested in order to improve the adhesion between CF and NCs. This process
could be carried out by coating the CF surface with a solution of TPU
nanocomposites combined with treated GNPs or hybrid nanofillers and prepared
by in-situ polymerisation so as to obtain MSCs with optimised properties.
➢ From the results of Mode-I ILFT, low-velocity impact and CAI tests of MSCs, it
is recommended to use a treated nanofiller or hybrid nanocomposites with a
loading of fillers below 1 wt.%. This proposal is built on the improvement in the
ILFT of MSCs based on the low loading of GNPs that was achieved during this
study.
➢ This study concentrated on using nanocomposites and MSCs with TPU with 70
wt.% HS as a matrix. A polymer blend of TPU with two different hard segments,
such as 60 wt.% and 90 wt.%, can also be used to produce NCs and MSCs. The
previous suggestion would help to obtain a polymer-blend matrix with high
mechanical properties; moreover, by incorporation with treated GNPs, the
interfacial adhesion of NCs and MSCs will also be improved.

261
References

[1] Y. Liu, "Nano-Reinforced Epoxy Resin For Carbon Fibre Fabric Composites," Doctor
of Philosophy in the Faculty of Engineering and Physical Science, University of
Manchester, 2016.
[2] T. J. Reinhart, S. T. Peters, Ed. Overview of Composite: Handbook of composites, 2nd
ed. London, England: Chapman & Hall`, 1998.
[3] Robert M. Jones, Mechanics of composite materials, 2nd ed. Unites states of
America: Taylor & Francis, 1999.
[4] P. Morgan, Polymer matrices for carbon fiber composites.In: Carbon fiber and their
composites. New York: Taylor and Francis Group, 2005.
[5] S. U. Khan and J.-K. Kim, "Impact and Delamination Failure of Multiscale Carbon
Nanotube-Fiber Reinforced Polymer Composites: A Review," International Journal
of Aeronautical and Space Sciences, vol. 12, no. 2, pp. 115-133, 2011.
[6] T.-W. Shyr and Y.-H. Pan, "Impact resistance and damage characteristics of
composite laminates," Composite Structures, vol. 62, no. 2, pp. 193-203, 2003.
[7] W. J. Cantwell and J. Morton, "The impact resistance of composite materials, a
review," Composites, vol. 22, no. 5, 1991.
[8] E. Selver, "Tow Level Hybridisation for Damage Tolerant Composites," University of
Manchester, 2014.
[9] A. M. Dı´ez-Pascual et al., "Influence of carbon nanotubes on the thermal, electrical
and mechanical properties of poly(ether ether ketone)/glass fiber laminates," CA R
B O N, vol. 49, pp. 2817 –2833, 2011.
[10] R. Junid, "Multiscale Carbon Fibre Composites with Epoxy-Graphite Nanoplatelet
Matrices," Doctor of Philosophy (PhD), School of Materials, Faculty of Engineering
and Physical Sciences, Unversity of Manchester, 2016.
[11] V. Cecen, M. Sarikanat, Y. Seki, T. Govsa, H. Yildiz, and I. H. Tavman, "Polyester
composites reinforced with noncrimp stitched carbon fabrics: Mechanical
characterization of composites and investigation on the interaction between
polyester and carbon fiber," Journal of Applied Polymer Science, vol. 102, no. 5, pp.
4554-4564, 2006.
[12] Y. Ma, M. Ueda, T. Yokozeki, T. Sugahara, Y. Yang, and H. Hamada, "A comparative
study of the mechanical properties and failure behavior of carbon fiber/epoxy and
carbon fiber/polyamide 6 unidirectional composites," Composite Structures, vol.
160, pp. 89-99, 2017.
[13] Y. Ma, T. Yokozeki, M. Ueda, T. Sugahara, Y. Yang, and H. Hamada, "Effect of
polyurethane dispersion as surface treatment for carbon fabrics on mechanical
properties of carbon/Nylon composites," Composites Science and Technology, vol.
151, pp. 268-281, 2017.
[14] Y. Ma, Y. Yang, T. Sugahara, and H. Hamada, "A study on the failure behavior and
mechanical properties of unidirectional fiber reinforced thermosetting and
thermoplastic composites," Composites Part B: Engineering, vol. 99, pp. 162-172,
2016.
[15] H. J. Zo, S. H. Joo, T. Kim, P. S. Seo, J. H. Kim, and J. S. Park, "Enhanced Mechanical
and Thermal Properties of Carbon Fiber Composites with Polyamide and
Thermoplastic Polyurethane Blends," Fibers and Polymers, vol. 15, no. 5, pp. 1071-
1077, 2014.
[16] S. A. Abdullah, A. Iqbal, and L. Frormann, "Melt mixing of carbon fibers and carbon
nanotubes incorporated polyurethanes," Journal of Applied Polymer Science, vol.
110, no. 1, pp. 196-202, 2008.

262
[17] Q. L. Shuai Jiang, Junwei W., Zhenglong He, Yuhua Z., Maoqing K., "Multiscale
graphene oxide–carbon fiber reinforcements for advanced polyurethane
composites," Composites : Part A, vol. 87, pp. 1-9, 2016.
[18] G. Zhao, T. Wang, and Q. Wang, "Surface modification of carbon fiber and its
effects on the mechanical and tribological properties of the polyurethane
composites," Polymer Composites, vol. 32, no. 11, pp. 1726-1733, 2011.
[19] A. M. Díez-Pascual, J. M. González-Domínguez, M. Teresa Martínez, and M. A.
Gómez-Fatou, "Poly(ether ether ketone)-based hierarchical composites for
tribological applications," Chemical Engineering Journal, vol. 218, pp. 285-294,
2013.
[20] Y. Li, J. Zhu, S. Wei, J. Ryu, L. Sun, and Z. Guo, "Poly(propylene)/Graphene
Nanoplatelet Nanocomposites: Melt Rheological Behavior and Thermal, Electrical,
and Electronic Properties," Macromolecular Chemistry and Physics, vol. 212, no. 18,
pp. 1951-1959, 2011.
[21] K. NA., "Materials science: carbon sheet solutions," Nature, vol. 442, pp. 254-5,
2006.
[22] B. Li and W.-H. Zhong, "Review on polymer/graphite nanoplatelet
nanocomposites," Journal of Materials Science, vol. 46, no. 17, pp. 5595-5614,
2011.
[23] K. Onyu et al., "Evaluation of the Possibility for Using Polypropylene/Graphene
Composite as Bipolar Plate Material Instead of Polypropylene/Graphite
Composite," KMUTNB International Journal of Applied Science and Technology, pp.
1-13, 2016.
[24] B. W. Chieng, N. A. Ibrahim, W. M. Yunus, M. Z. Hussein, and V. S. Giita Silverajah,
"Graphene nanoplatelets as novel reinforcement filler in poly(lactic
acid)/epoxidized palm oil green nanocomposites: mechanical properties," Int J Mol
Sci, vol. 13, no. 9, pp. 10920-34, 2012.
[25] M. Sabzi, L. Jiang, F. Liu, I. Ghasemi, and M. Atai, "Graphene nanoplatelets as
poly(lactic acid) modifier: linear rheological behavior and electrical conductivity,"
Journal of Materials Chemistry A, vol. 1, no. 28, p. 8253, 2013.
[26] J. Suh and D. Bae, "Mechanical properties of polytetrafluoroethylene composites
reinforced with graphene nanoplatelets by solid-state processing," Composites Part
B: Engineering, vol. 95, pp. 317-323, 2016.
[27] S. Chandrasekaran, C. Seidel, and K. Schulte, "Preparation and characterization of
graphite nano-platelet (GNP)/epoxy nano-composite: Mechanical, electrical and
thermal properties," European Polymer Journal, vol. 49, no. 12, pp. 3878-3888,
2013.
[28] Y. Geng, S. J. Wang, and J. K. Kim, "Preparation of graphite nanoplatelets and
graphene sheets," J Colloid Interface Sci, vol. 336, no. 2, pp. 592-8, Aug 15 2009.
[29] P. Pokharel, S. H. Lee, and D. S. Lee, "Thermal, Mechanical, and Electrical Properties
of Graphene Nanoplatelet/Graphene Oxide/Polyurethane Hybrid Nanocomposite,"
Journal of Nanoscience and Nanotechnology, vol. 15, no. 1, pp. 211-214, 2015.
[30] Carbon Fiber Composites Market 2019 Global Leading Players, Industry Updates,
Future Growth, Business Prospects, Developments and Investments Forecast to
2024 [Online]. Available: https://www.reuters.com/brandfeatures/venture-
capital/article?id=103638
[31] K. Pulidindi and H. Pandey. (2016). Carbon Fiber Composites Market Available:
https://www.gminsights.com/industry-analysis/carbon-fibre-composites-market
[32] T. Kraus and M. Kühnel, "The Global CRP Market," 6 October 2014 2014, Available:
https://eucia.eu › userfiles › files › 20141008_market_report_grpcrp.

263
[33] L. A. Pilato and J. Michno, Advanced composite materials, 1st ed. New York:
Springer-Verlag Berlin Heidelberg, 1994.
[34] P. K. Mallick, Fibre-Reinforced composite, 3rd ed. Taylor and Fracis Group, LLC,
2008.
[35] I. Y. Chang and J. K. Lees, "Recent Development in Thermoplastic Composites: A
Review of Matrix Systems and Processing Methods," Journa of thermoplastic
composite materials, vol. 1, no. 3, pp. 277-296, 1988.
[36] L. W. McKeen, The Effect of Temperature and Other Factors on plastic and
Elastomers, Second ed. Norwich, NY, USA: William Andrew Inc., 2008.
[37] R. J. Crawford, Plastics Engineering, 3rd ed. Butterworth-Heineman, 2002.
[38] J. G. Drobny, Handbook of Thermoplastic Elastomers:Chapter one, 2nd ed. United
state: Elsevier Inc., 2014.
[39] G. Holden, N. R. Legge, R. Quirk, and H. E. Schroeder, Thermoplastic Elastomers,
2nd ed. Munich: Hanser Publishers, 1996.
[40] M. Dollausen, G. Oretel, Ed. In Polyurethane Handbook. Hanser: Munich, 1985.
[41] G. Woods, "The ICI Polyurethane Book," New York: ICI Polyurethanes and John
Wiley, 1990.
[42] C. P. Pader, C. A. Harper, Ed. Thermoplastic Elastomers : Handbook of Plastic,
Elastomers, and Composites, 3rd ed. New York: McGraw-Hill Companies, 1996.
[43] Z. S. Petrovic and J. Ferguson, "Polyurethane Elastomers," Prog. Polym. Sci, vol. 16,
pp. 695-836, 1991.
[44] M. Szycher, Szycher's Handbook of Polyurethanes, 2nd ed. 300 Boca Raton, FL:
Taylor & Francis Group, LLC, 2013.
[45] C. Prisacariu, Polyurethane Elastomers From Morphology to Mechanical Aspects.
New York: Springer-Verlag/Wien, 2011.
[46] M. A. Albozahid, "Design of Novel High Modulus TPUs for NanoComposite
Applications," PhD, Faculty of Science and Engineering, University of Manchester,
2018.
[47] Z. K. M. Al-Obad, "Designing PU Resins for Fibre Composite Applications," Doctor of
Philosophy, Faculty of Science and Engineering, University of Manchester, 2017.
[48] H. Kim, Y. Miura, and C. W. Macosko, "Graphene/Polyurethane Nanocomposites for
Improved Gas Barrier and Electrical Conductivity," Chemistry of Materials, vol. 22,
no. 11, pp. 3441-3450, 2010.
[49] F. Askari, M. Barikani, M. Barmar, and P. Shokrollahi, "Polyurethane/amino-grafted
multiwalled carbon nanotube nanocomposites: Microstructure, thermal,
mechanical, and rheological properties," Journal of Applied Polymer Science, vol.
134, no. 4, 2017.
[50] H. Xia and M. Song, "Preparation and characterisation of polyurethane grafted
single-walled carbon nanotubes and derived polyurethane nanocomposites,"
Journal of Materials Chemistry, vol. 16, no. 19, p. 1843, 2006.
[51] Sanchez-Adsuar, M. Pastor-Blas, and J. M. Martin-Martinez, "Properties of
Polyurethane Elastomers with Different Hard/Soft Segment Ratio," The Journal of
Adhesion, vol. 67, p. 327, 1998.
[52] A. A. Tsiotas, "The role of the chain extender on the phase behaviour and
morphology of high hard block content thermoplastic
polyurethanes:Thermodynamics-Structures-Properties," PhD, School of materials,
Manchester, 2012.
[53] T. W. Son, D. W. Lee, and S. K. Lim, "Thermal and Phase Behavior of Polyurethane
Based on Chain Extender,2,2-Bis-[4-(2-hydroxyethoxy)phenyl]propane," Polymer
Journal, vol. 31, no. 7, pp. 563-568, 1999.

264
[54] D. K. Chattopadhyay and D. C. Webster, "Thermal stability and flame retardancy of
polyurethanes," Progress in Polymer Science, vol. 34, no. 10, pp. 1068-1133, 2009.
[55] R. W. Seymour and S. L. Cooper, "Thermal Analysis of Polyurethane Block
Polymers," Macromolecules, vol. 6, no. 1, pp. 48-53, 1973.
[56] K. D. Kavlock, T. W. Pechar, J. O. Hollinger, S. A. Guelcher, and A. S. Goldstein,
"Synthesis and characterization of segmented poly(esterurethane urea) elastomers
for bone tissue engineering," Acta Biomater, vol. 3, no. 4, pp. 475-84, Jul 2007.
[57] H. Liu et al., "Preparation and characterization of aliphatic polyurethane and
hydroxyapatite composite scaffold," Journal of Applied Polymer Science, vol. 112,
no. 5, pp. 2968-2975, 2009.
[58] G. Sankar and N. Yan, "Synthesis and Deblocking Studies of Low Temperature Heat-
Curable Blocked Polymeric Methylene Diphenyl Diisocyanates," Journal of
Macromolecular Science, Part A, vol. 52, no. 1, pp. 47-55, 2014.
[59] P. A. Gunatillake, D. J. Martin, G. F. Meijs, S. J. McCarthy, and R. Adhikari,
"Designing Biostable Polyurethane Elastomers for Biomedical Implants," Csiro
Publishing, pp. 545–557, 2003.
[60] P. Vermette, H. J. Griesser, G. Laroche, and R. Guidoin, Biomedical Applications of
Polyurethanes, Tissue Engineering Intelligence Unit 6 Georgetown, Texas, U.S.A.:
EUREKAH.COM, 2001.
[61] A. TO, C. IS, J. HM, and C. K, "Thermal and mechanical-properties of thermoplastic
polyurethane elastomers from different polymerization methods.," Polmer
international, vol. 31, pp. 329–333, 1993.
[62] Y. He, D. Xie, and X. Zhang, "The structure, microplase-separated morphology, and
property of polyurethane and polyureas," J. Mater. Sci., vol. 49, pp. 7339-7352,
2014.
[63] S. L. Cooper and J. Guan, Advances in Polyurethane Biomaterials:Woodhead
Publishing Series in Biomaterials:Number 108. Langford Lane, Kidlington, UK:
Elsevier Ltd., 2016.
[64] R. W. Seymour, G. M. Estes, and S. L. Cooper, "Infrared Studies of Segmented
Polyurethan Elastomers. I. Hydrogen Bonding," Macromolecules, vol. 3, no. 5, p.
579, 1970.
[65] K. Knutson, and Lyman, D. J., "In Biomaterials: Interfacial Phenomena and
Applications(Edited by S. L. Cooper and N. A. Peppas)," Adv.Chem.Ser., vol. 199, no.
109, 1982.
[66] J. Bandekar and S. Klima, "FT-IR spectroscopic studies of polyurethanes--Part II. Ab
initio quantum chemical studies of the relative strengths of "carbonyl" and "ether"
hydrogen-bonds in polyurethanes," Specrochimica Acta, vol. 48A, no. 10, pp. 1363-
1370, 1992.
[67] Z. Ren, D. Ma, and X. Yang, "H-bond and conformations of donors and acceptors in
model polyether based poyurethanes," Polymer Journal, vol. 44, pp. 6419-6425,
2003.
[68] C. Hepburn, Polyurethane Elastomers, 2nd ed. London: Elsevier Applied Science,
1992.
[69] D. J. Martin, G. F. Meijs, P. A. Gunatillake, S. J. Mccarthy, and G. M. Renwick, "The
Effect of Average Soft Segment Length on Morphology and Properties of a Series of
Polyurethane Elastomers. II. SAXS-DSC Annealing Study," Journal of Applied Polymer
Science, vol. 64, no. 4, pp. 803-817, 1997.
[70] K. Jirakittidul, "Structure-Property Relationships in Polyurethane-Carbon Particle
Nanocomposites," PhD, School of Materials,PhD, Manchester University, 2013.

265
[71] A. Saiani, W. A. Daunch, H. Verbeke, J. W. Leengslag, and J. S. Higgins, "Origin of
Multiple Melting Endotherms in a High Hard Block Content Polyurethane. 1.
Thermodynamic Investigation," Macromolecules, vol. 34, pp. 9059-9068, 2001.
[72] A. Saiani, A. Novak, G. Eeckhaut, L. Rodier, J. W. Leenslag, and J. S. Higgins, "Origin
of Multiple Melting Endotherms in a High Hard Block Content Polyurethane: Effect
of Annealing Temperature," Macromolecules, vol. 40, pp. 7252-7262, 2007.
[73] A. Saiani, G. Rochas, G. Eeckhaut, W. A. Daunch, J. W. Leenslag, and J. S. Higgins,
"Origin of Multiple Melting Endotherms in a High Hard Block Content Polyurethane.
2. Structural Investigation," Macromolecules, vol. 37, pp. 1411-1421, 2003.
[74] J. T. Koberstein and T. P. Russell, "Simultaneous SAXS-DSC Study of Multiple
Endothermic Behavior in Polyether-Based Polyurethane Block Copolymers,"
Macromolecules vol. 19, pp. 714-720, 1986.
[75] P. E. Gibson, M. A. Vallence, and S. L. Cooper, In Development in Block Copolymers.
London: Goodman, I., Ed.; Applied Science Series; Elsevier, 1982.
[76] Y. Li, T. Gao, J. Liu, R. Linliu, C. R. Deeper, and B. Chu, "Multiphase Structure of a
Segmented Polyurethane: Effect of Temperature and Annealing.," Macromolecules,
vol. 25, pp. 7365-7372, 1992.
[77] J. W. C. Van Bogart, P. E. Gibson, and S. L., "Structure-Property Relationships in
Polycaprolactone-Polyurethanes," J. of Polymer Science: Polymer Physics Edition,
vol. 21, pp. 65-95, 1983.
[78] B. S. Lee, B. C. Chun, Y.-C. Chung, K. I. Sul, and J. W. Cho§, "Structure and
Thermomechanical Properties of Polyurethane Block Copolymers with Shape
Memory Effect," Macromolecules, vol. 34, pp. 6431-6437, 2001.
[79] R. A. Nallicheri and M. F. Rubner, "Thermal and Mechanical Properties of
Polyurethane-Diacetylene Segmented Copolymers. 1. Molecular Weight and
Annealing Effects," Macromolecules, vol. 23, pp. 1005-1016, 1990.
[80] P. R. Laity et al., "A 2-dimensional small-angle X-ray scattering study of the
microphase-separated morphology exhibited by thermoplastic polyurethanes and
its response to deformation," Polymer, vol. 45, no. 15, pp. 5215-5232, 2004.
[81] C. Prisacariu, R. H. Olley, A. A. Caraculacu, D. C. Bassett, and C. Martin, "The effect
of hard segment ordering in copolyurethane elastomers obtained by using
simultaneously two types of diisocyanates," Polymer, vol. 44, no. 18, pp. 5407-
5421, 2003.
[82] Y. Yanagihara, N. Osaka, S. Iimori, S. Murayama, and H. Saito, "Relationship
between modulus and structure of annealed thermoplastic polyurethane,"
Materials Today Communications, vol. 2, pp. e9-e15, 2015.
[83] J. W. Zondlo, p. mukhopadhyay and r. K. Gupta, Eds. Graphite: Structure, Properties,
and Applications in Graphite, Graphene and their Polymer Nanocomposites. Broken
Sound Parkway NW, Boca Raton: Taylor & Francis Group, LLC, 2013.
[84] Y. Geng, J. Li, and J. K. Kim, S. C. Tjong and Y.-W. Mai, Eds. "Synthesis and electrical
conducting behavior of graphite nanoplatelet/polymer nanocomposites" In Physical
properties and applications of polymer nanocomposites Woodhead Publishing
Limited, 2010.
[85] R. Sengupta, M. Bhattacharya, S. Bandyopadhyay, and A. K. Bhowmick, "A review
on the mechanical and electrical properties of graphite and modified graphite
reinforced polymer composites," Progress in Polymer Science, vol. 36, no. 5, pp.
638-670, 2011.
[86] M. Song and D. Cai, V. Mittal, Ed. "Graphene Functionalization: A Review" in
Polymer–Graphene Nanocomposites. Dorchester, UK: The Royal Society of
Chemistry, 2012.

266
[87] K. S. Novoselov et al., "Electric Field Effect in Atomically Thin Carbon Films,"
Science, vol. 306, pp. 666-669, 2004.
[88] A. K. Geim and K. S. Novoselov, "The rise of graphene," Nature Material vol. 6,
2007.
[89] L. I. Choi W, Seelaboyina R, Kang YS., "Synthesis of graphene and its applications: a
review," Critical Reviews in Solid State and Materials Sciences, vol. 35, pp. 52-71,
2010.
[90] A. V. Yakovlev, A. I. Finaenov, S. L. Zabud’kov, and E. V. Yakovleva, "Thermally
expanded graphite: Synthesis, properties, and prospects for use," Russian Journal of
Applied Chemistry, vol. 79, no. 11, pp. 1741-1751, 2006.
[91] S. Park and R. S. Ruoff, "Chemical methods for the production of graphenes," Nat
Nanotechnol, vol. 4, no. 4, pp. 217-24, Apr 2009.
[92] J. Du and H.-M. Cheng, "The Fabrication, Properties, and Uses of Graphene/Polymer
Composites," Macromolecular Chemistry and Physics, vol. 213, no. 10-11, pp. 1060-
1077, 2012.
[93] S. Kashi, R. K. Gupta, N. Kao, and S. N. Bhattacharya, "Viscoelastic properties and
physical gelation of poly (butylene adipate-co-terephthalate)/graphene
nanoplatelet nanocomposites at elevated temperatures," Polymer, vol. 101, pp.
347-357, 2016.
[94] H. Pang, L. Xu, D.-X. Yan, and Z.-M. Li, "Conductive polymer composites with
segregated structures," Progress in Polymer Science, vol. 39, no. 11, pp. 1908-1933,
2014.
[95] D. G. Papageorgiou, I. A. Kinloch, and R. J. Young, "Graphene/elastomer
nanocomposites," Carbon, vol. 95, pp. 460-484, 2015.
[96] E. Thostenson, C. Li, and T. Chou, "Nanocomposites in context," Composites Science
and Technology, vol. 65, no. 3-4, pp. 491-516, 2005.
[97] A. Duguay, "Exfoliated Graphite Nanoplatelet-Filled Impact Modified Polypropylene
Nanocomposites," Master University of Maine, 2011.
[98] M. Alexandre and P. Dubois, "Polymer-layered silicate nanocomposites:
preparation, properties and uses of a new class of materials," Materials Science and
Engineering, vol. 28, pp. 1-63, 2000.
[99] J. Vermant, S. Ceccia, M. K. Dolgovskij, P. L. Maffettone, and C. W. Macosko,
"Quantifying dispersion of layered nanocomposites via melt rheology," Journal of
Rheology, vol. 51, no. 3, pp. 429-450, 2007.
[100] T. G. Mezger, The rheology handbook: for users of rotational and oscillatory
rheometers, 2nd revised edition ed. Germany: Vincentz Network GmbH & Co. KG,
Hannover, 2006.
[101] S. Basu, M. Singhi, B. K. Satapathy, and M. Fahim, "Dielectric, electrical, and
rheological characterization of graphene-filled polystyrene nanocomposites,"
Polymer Composites, vol. 34, no. 12, pp. 2082-2093, 2013.
[102] M. M. Shokrieh, M. Esmkhani, H. R. Shahverdi, and F. Vahedi, "Effect of Graphene
Nanosheets (GNS) and Graphite Nanoplatelets (GNP) on the Mechanical Properties
of Epoxy Nanocomposites," Science of Advanced Materials, vol. 5, no. 3, pp. 260-
266, 2013.
[103] H.-B. Zhang, W.-G. Zheng, Q. Yan, Z.-G. Jiang, and Z.-Z. Yu, "The effect of surface
chemistry of graphene on rheological and electrical properties of
polymethylmethacrylate composites," Carbon, vol. 50, no. 14, pp. 5117-5125, 2012.
[104] C. McClory, T. McNally, M. Baxendale, P. Pötschke, W. Blau, and M. Ruether,
"Electrical and rheological percolation of PMMA/MWCNT nanocomposites as a
function of CNT geometry and functionality," European Polymer Journal, vol. 46, no.
5, pp. 854-868, 2010.

267
[105] P. PoÈtschke, T. D. Fornes, and D. R. Paul, "Rheological behavior of multiwalled
carbon nanotube/polycarbonate composites," polymer, vol. 43, pp. 3247-3255,
2002.
[106] H. Kim and C. W. Macosko, "Processing-property relationships of
polycarbonate/graphene composites," Polymer, vol. 50, no. 15, pp. 3797-3809,
2009.
[107] J. T. Choi et al., "Functionalized graphene sheet/polyurethane nanocomposites:
Effect of particle size on physical properties," Macromolecular Research, vol. 19, no.
8, pp. 809-814, 2011.
[108] Y. Gao, O. T. Picot, E. Bilotti, and T. Peijs, "Influence of filler size on the properties of
poly(lactic acid) (PLA)/graphene nanoplatelet (GNP) nanocomposites," European
Polymer Journal, vol. 86, pp. 117-131, 2017.
[109] H. S. Kim, H. S. Bae, J. Yu, and S. Y. Kim, "Thermal conductivity of polymer
composites with the geometrical characteristics of graphene nanoplatelets," Sci
Rep, vol. 6, p. 26825, May 25 2016.
[110] D. Dean, A. M. Obore, S. Richmond, and E. Nyairo, "Multiscale fiber-reinforced
nanocomposites: Synthesis, processing and properties," Composites Science and
Technology, vol. 66, no. 13, pp. 2135-2142, 2006.
[111] E. Kandare et al., "Improving the through-thickness thermal and electrical
conductivity of carbon fibre/epoxy laminates by exploiting synergy between
graphene and silver nano-inclusions," Composites Part A: Applied Science and
Manufacturing, vol. 69, pp. 72-82, 2015.
[112] J.-h. Han et al., "The combination of carbon nanotube buckypaper and insulating
adhesive for lightning strike protection of the carbon fiber/epoxy laminates,"
Carbon, vol. 94, pp. 101-113, 2015.
[113] M. Gagné and D. Therriault, "Lightning strike protection of composites," Progress in
Aerospace Sciences, vol. 64, pp. 1-16, 2014.
[114] A. M. Díez-Pascual, M. Naffakh, C. Marco, M. A. Gómez-Fatou, and G. J. Ellis,
"Multiscale fiber-reinforced thermoplastic composites incorporating carbon
nanotubes: A review," Current Opinion in Solid State and Materials Science, vol. 18,
no. 2, pp. 62-80, 2014.
[115] W. Qin, F. Vautard, L. T. Drzal, and J. Yu, "Mechanical and electrical properties of
carbon fiber composites with incorporation of graphene nanoplatelets at the fiber–
matrix interphase," Composites Part B: Engineering, vol. 69, pp. 335-341, 2015.
[116] E. Fitzer, F. J.L., B. C.A., B. R.T.K., and H. K.J., Eds. In Carbon Fibres Filaments and
Composites. Dordrecht: Kluwer Academic, 1990.
[117] S.-J. Park, R. Hull et al., Eds. Carbon fibres (Springer Series in Materials Science).
London: Springer Science and Business Media, 2015.
[118] W. S. Smith, Engineered materials handbook-Vol.1 (ASM International. Ohio, 1987.
[119] S. Chand, "Review carbon fibers for composites.," journal of Materials Science, vol.
35, pp. 1303-1313, 2000.
[120] J.-K. Kim and Y.-W. Mai, Engineered interfaces in fibre reinforced composites,
chapter 5, . New York: Elsevier, 1998.
[121] F. H. Gojny, M. H. G. Wichmann, B. Fiedler, W. Bauhofer, and K. Schulte, "Influence
of nano-modification on the mechanical and electrical properties of conventional
fibre-reinforced composites," Composites Part A: Applied Science and
Manufacturing, vol. 36, no. 11, pp. 1525-1535, 2005.
[122] V. Kostopoulos, A. Baltopoulos, P. Karapappas, A. Vavouliotis, and A. Paipetis,
"Impact and after-impact properties of carbon fibre reinforced composites
enhanced with multi-wall carbon nanotubes," Composites Science and Technology,
vol. 70, no. 4, pp. 553-563, 2010.

268
[123] B. Ashrafi et al., "Processing and properties of PEEK/glass fiber laminates: Effect of
addition of single-walled carbon nanotubes," Composites Part A: Applied Science
and Manufacturing, vol. 43, no. 8, pp. 1267-1279, 2012.
[124] Z. Shen, S. Bateman, D. Wu, P. McMahon, M. Dellolio, and J. Gotama, "The effects
of carbon nanotubes on mechanical and thermal properties of woven glass fibre
reinforced polyamide-6 nanocomposites," Composites Science and Technology, vol.
69, no. 2, pp. 239-244, 2009.
[125] A. Godara et al., "Influence of carbon nanotube reinforcement on the processing
and the mechanical behaviour of carbon fiber/epoxy composites," Carbon, vol. 47,
no. 12, pp. 2914-2923, 2009.
[126] K.-Y. Kim, L. Ye, and K.-M. Phoa, "Interlaminar Fracture Toughness of CF/PEI and
GF/PEI Composites at Elevated Temperatures," Applied Composite Materials, vol.
11, pp. 173–190, 2004.
[127] E. Borowski, E. Soliman, U. F. Kandil, and M. R. Taha, "Interlaminar Fracture
Toughness of CFRP Laminates Incorporating Multi-Walled Carbon Nanotubes,"
Polymers, vol. 7, no. 6, pp. 1020-1045, 2015.
[128] T. D. Fornes and D. R. Paul, "Modeling properties of nylon 6/clay nanocomposites
using composite theories," Polymer, vol. 44, no. 17, pp. 4993-5013, 2003.
[129] E. Bekyarova et al., "Multiscale Carbon Nanotube-Carbon Fiber Reinforcement for
Advanced Epoxy Composites," Langmuir vol. 23, pp. 3970-3974, 2007. American
Chemical Society
[130] M. Kim, Y.-B. Park, O. I. Okoli, and C. Zhang, "Processing, characterization, and
modeling of carbon nanotube-reinforced multiscale composites," Composites
Science and Technology, vol. 69, no. 3-4, pp. 335-342, 2009.
[131] A. F. Ávila, L. G. Z. d. O. Peixoto, A. Silva Neto, J. d. Á. Junior, and M. G. R. Carvalho,
"Bending Investigation on Carbon Fiber/Epoxy Composites NanoModified by
Graphene," J. Brazilian Soc. Mech. Sci. Eng, vol. 34, no. 3, pp. 269-275, 2012.
[132] A. F. Ávila, M. I. Yoshida, M. G. R. Carvalho, E. C. Dias, and J. de Ávila, "An
investigation on post-fire behavior of hybrid nanocomposites under bending loads,"
Composites Part B: Engineering, vol. 41, no. 5, pp. 380-387, 2010.
[133] M. M. Rahman et al., "Effect of NH2-MWCNTs on crosslink density of epoxy matrix
and ILSS properties of e-glass/epoxy composites," Composite Structures, vol. 95, pp.
213-221, 2013.
[134] W. Liu, L. Li, S. Zhang, F. Yang, and R. Wang1, "Mechanical Properties of Carbon
Nanotube/Carbon Fiber Reinforced Thermoplastic Polymer Composite," Polymer
Composites, 2015. Society of Plastics Engineers
[135] P. Robinson and J. M. Hodgkinson, J. M. Hodgkinson, Ed. Interlaminar fracture
toughness in "Mechanical testing of advanced fibre composites" Cambridge,
England: Woodhead Publishing Limited, 2000
[136] E. Oterkus, C. Diyaroglu, D. De Meo, and G. Allegri, J. Graham-Jones and J.
Summerscales, Eds. Chapter 4- Fracture modes, damage tolerance and failure
mitigation in marine composites (in Marine Applications of Advanced Fibre-
Reinforced Composites). Eds. Elsevier, 2016.
[137] P. Wang, W. Liu, X. Zhang, X. Lu, and J. Yang, "Enhanced fracture toughness of
carbon fabric/epoxy laminates with pristine and functionalized stacked-cup carbon
nanofibers," Engineering Fracture Mechanics, vol. 148, pp. 73-81, 2015.
[138] T. Yokozeki, Y. Iwahori, S. Ishiwata, and K. Enomoto, "Mechanical properties of
CFRP laminates manufactured from unidirectional prepregs using CSCNT-dispersed
epoxy," Composites Part A: Applied Science and Manufacturing, vol. 38, no. 10, pp.
2121-2130, 2007.

269
[139] M. O. W. Richardson and M. J. Wisheart, "Review of low-velocity impact properties
of composite materials," Composites Part A: Applied Science and Manufacturing,
vol. 27, no. 12, pp. 1123-1131, 1996.
[140] H. Ulus, Ö. S. Şahin, and A. Avcı, "Enhancement of flexural and shear properties of
carbon fiber/epoxy hybrid nanocomposites by boron nitride nano particles and
carbon nano tube modification," Fibers and Polymers, vol. 16, no. 12, pp. 2627-
2635, 2016.
[141] K. Iqbal, S.-U. Khan, A. Munir, and J.-K. Kim, "Impact damage resistance of CFRP
with nanoclay-filled epoxy matrix," Composites Science and Technology, vol. 69, no.
11-12, pp. 1949-1957, 2009.
[142] T. H. Mahdi, M. E. Islam, M. V. Hosur, and S. Jeelani, "Low-velocity impact
performance of carbon fiber-reinforced plastics modified with carbon nanotube,
nanoclay and hybrid nanoparticles," Journal of Reinforced Plastics and Composites,
vol. 36, no. 9, pp. 696-713, 2017.
[143] F. Inam, D. W. Y. Wong, M. Kuwata, and T. Peijs, "Multiscale Hybrid Micro-
Nanocomposites Based on Carbon Nanotubes and Carbon Fibers," Journal of
Nanomaterials, vol. 2010, pp. 1-12, 2010.
[144] V. Eskizeybek, H. Ulus, H. B. Kaybal, Ö. S. Şahin, and A. Avcı, "Static and dynamic
mechanical responses of CaCO3 nanoparticle modified epoxy/carbon fiber
nanocomposites," Composites Part B: Engineering, vol. 140, pp. 223-231, 2018.
[145] H. Ulus, T. Üstün, Ö. S. Şahin, S. E. Karabulut, V. Eskizeybek, and A. Avcı, "Low-
velocity impact behavior of carbon fiber/epoxy multiscale hybrid nanocomposites
reinforced with multiwalled carbon nanotubes and boron nitride nanoplates,"
Journal of Composite Materials, vol. 50, no. 6, pp. 761-770, 2015.
[146] A. Markov, B. Fiedler, and K. Schulte, "Electrical conductivity of carbon black/fibres
filled glass-fibre-reinforced thermoplastic composites," Composites Part A: Applied
Science and Manufacturing, vol. 37, no. 9, pp. 1390-1395, 2006.
[147] K. A. Imran and K. N. Shivakumar, "Graphene-modified carbon/epoxy
nanocomposites: Electrical, thermal and mechanical properties," Journal of
Composite Materials, vol. 53, no. 1, pp. 93-106, 2018.
[148] Y. Li, H. Zhang, Z. Huang, E. Bilotti, and T. Peijs, "Graphite Nanoplatelet Modified
Epoxy Resin for Carbon Fibre Reinforced Plastics with Enhanced Properties," Journal
of Nanomaterials, vol. 2017, pp. 1-10, 2017.
[149] (2018). Sigma Aldrich Company, Sigma Aldrich, UK Available:
www.sigmaaldrich.com/united-kingdom
[150] X. S. XG Sciences Company, Inc,. (2018). xGnP® Graphene Nanoplatelets – Grade M.
Available: https://xgsciences.com/wp-content/uploads/2017/11/xGnP-M.-
MD00003.-2018-1.pdf
[151] Sigmatex Materials innovatio. (2019). SC5171295 Sigmatex Product Data Sheet.
Available: http://www.sigmatex.com/
[152] J. N. Coleman, U. Khan, and Y. K. Gun'ko, "Mechanical Reinforcement of Polymers
Using Carbon Nanotubes," Advanced Materials, vol. 18, no. 6, pp. 689-706, 2006.
[153] V. Mittal, Polymer-graphene nanocomposites. Thomas Graham House, Science
Park, Milton Road, Cambridge CB4 0WF, UK: The Royal Society of Chemistry, 2012.
[154] T. Y. Tsai, C. H. Li, C. H. Chang, W. H. Cheng, C. L. Hwang, and R. J. Wu, "Preparation
of Exfoliated Polyester/Clay Nanocomposites," Advanced Materials, vol. 17, no. 14,
pp. 1769-1773, 2005.
[155] BSI - BS EN ISO 527-2 (Plastic - Determination of tensile properties part 2: Test
conditions for moulding and extrusion plastics), 2012.
[156] W.-F. Su, Polymer Size and Polymer Solutions, in Principles of Polymer Design and
Synthesis. Berlin Heidelberg: Springer-Verlag, 2013.

270
[157] G. Odian, Principles of Polymerization, Fourth Edi ed. New Jersey: John Wiley &
Sons, Inc., 2004.
[158] R. P. Redman, J. M. Buist, Ed. Developments in Polyurethane Elastomers, in
Developments in Polyurethane. Applied Science Publishers, 1978.
[159] A. e. M. Striegel, W. W. Yau, J. J. Kirkland, and D. D. Bly, second, Ed. Modern Size-
Exclusion Liquid Chromatography. Hoboken, New Jersey: John Wiley & Sons, Inc.,
2009.
[160] V. Bershtein and V. Egorov, Differential Scanning Calorimetry of Polymers: Physics,
Chemistry, Analysis, Technology. New York: Ellis Horwood, 1994.
[161] P. G. Laye, P. J. Haines, Ed. Differential Thermal Analysis and Differential Scanning
Calorimetry, in Principles of Thermal Analysis and Calorimetry. Royal Society of
Chemistry, 2002.
[162] Y. Leng, Materials Characterization Introduction to Microscopic and Spectroscopic
Methods. John Wiley & Sons, 2009.
[163] T. Hatakeyama and F. X. Quinn, Thermal analysis: Fundamentals and applications to
polymer science, second ed. Baffins Lane, Chichester,England: John Wiley & Sons
Ltd, 1999.
[164] B. H. Stuart, Polymer Analysis.s.I. John Wiley And Sons, LTD, 2003.
[165] TA Instruments Dynamic Mechanical Analyzer. (2018, Online). Available:
http://www.tainstruments.com/pdf/literature/TA284.pdf.
[166] D. M. Price, "Thermomechanical, Dynamic Mechanical and Dielectric Methods in
Principles of Thermal Analysis and Calorimetry," P. J. Haines, Ed. Royal Society of
Chemistry, 2002, pp. 94-128.
[167] L. E. Nielsen and R. F. Landel, "Mechanical Properties of Polymers and
Composites.," vol. 90, ed: CRC Press., 1994.
[168] Joseph D. Menczel and R. B. Prime., Thermal analysis of polymers. Hoboken, New
Jersey: John Wiley & Sons, INC, 2009.
[169] G. R. Heal, P. J. Haines, Ed. Thermogravimetry and Derivative Thermogravimetry, in
Principles of Thermal Analysis and Calorimetry. Royal Society of Chemistry., 2002.
[170] R. E. Wetton, B. J. Hunt and M. I. James, Eds. Thermal Analysis, in Polymer
Characterisation. Springer, 1993.
[171] C 518 – 04, Standard Test Method for Steady-State Thermal Transmission Properties
by Means of the Heat Flow Meter Apparatus, 2004.
[172] D. K. Bowen and B. K. Tanner, High Resolution X-ray Diffractometry and
Topography, 2nd ed. UK: Taylor & Francis Ltd, 2005.
[173] Anjali Bishnoi, Sunil Kumar, and N. Joshi., R. T. Sabu Thomas, Ajesh K. Zachariah,
Raghvendra Kumar Mishra, Ed. 'Wide-Angle X-ray Diffraction (WXRD): Technique for
Characterization of Nanomaterials and Polymer Nanocomposites' in Microscopy
Methods in Nanomaterials Characterization. Elsevier Inc., 2017.
[174] P. V. d. Heide, X-Ray Photoelectron Spectroscopy An Introduction to Principles and
Practices. Hoboken, New Jersey: John Wiley and Sons, Inc., 2012.
[175] A. W. Wren, F. R. Laffir, N. P. Mellot, and M. R. Towler1, J. M. Wagner, Ed. X-ray
Photoelectrons Spectroscopy: Studies From Industrial and Bioactive Glass to
Biomaterials , In X-Ray Photoelectron Spectroscopy. New York: Nova Science
Publishers, Inc, 2011.
[176] Dan Campbell, R.A. Pethrick, and J. R. White., Polymer Characterization: Physical
Techniques, 2nd ed. Taylor & Francis Group, 2000.
[177] D. B. Williams and C. B. Carter, Transmission Electron Microscopy A Textbook for
Materials Science. Springer Science and Business Media, 2009.
[178] R. F. Egerton, Physical Principles of Electron Microscopy: An Introduction to TEM,
SEM, and AEM. Springer Science & Business Media, 2006.

271
[179] C. D. Han, Rheology and Processing of Polymeric Materials. Oxford University Press,
Inc., 2007.
[180] J.-C. Majesté, S. Thomas, R. Muller, and J. Abraham, Eds. Rheology and processing
of polymer nanocomposites: Theory, practice and new challenges. Rheology and
Processing of Polymer Nanocomposites. John Wiley & Sons, Inc.,, 2016, pp.
Hoboken, New Jersey.
[181] R. P. Chhabra and J. F. Richardson, Non-Newtonian Flow and Applied Rheology.
Engineering Applications, 2nd ed. Amsterdam; Boston: Butterworth-
Heinemann/Elsevier, 2008.
[182] R. W. Whorlow, Rheological techniques. Ellis Horwood Limited, 1992.
[183] C. W. Macosko, Rheology Principles, Measurements and Application. Canada: Wiley-
VCH, Inc., 1994.
[184] "ASTM E1461 − 13, Standard Test Method for Thermal Diffusivity by the Flash
Method, ASTM E1461 − 13" 2013.
[185] W. J. Parker, R. J. Jenkins, C. P. Butler, and G. L. Abbott, "Flash Method of
Determining Thermal Diffusivity, Heat Capacity, and Thermal Conductivity," Journal
of Applied Physics, vol. 32, no. 9, pp. 1679-1684, 1961.
[186] M. Gresil, Z. Wang, Q. A. Poutrel, and C. Soutis, "Thermal Diffusivity Mapping of
Graphene Based Polymer Nanocomposites," Sci Rep, vol. 7, no. 1, p. 5536, Jul 17
2017.
[187] B. I.-. Plastics — Differential scanning calorimetry (DSC) — Part 4: Determination of
specific heat capacity, 2005.
[188] C. Schick, "Differential scanning calorimetry (DSC) of semicrystalline polymers,"
Anal Bioanal Chem, vol. 395, no. 6, pp. 1589-611, Nov 2009.
[189] R. Smith, D. Bruce, L. D. Jones, A. Marriott, L. P. Scudder, and S. Willsher, D. O.
Thompson and D. Chimenti, Eds. Ultrasonic C-Scan Standardization for Polymer-
Matrix Composites-Acoustic Considerations, in Review of Progress in Quantitative
Nondestructive Evaluation. New York: Plenum Press, 1998.
[190] P. Vaara and J. Leinonen, Technology Survey on NDT of Carbon-fiber Composites.
Kemi-Tornio University of Applied Sciences, 2012.
[191] T. L. Ajay Kapadia, "Non Destructive Testing of Composite Materials," National
Composite Network.
[192] Standard Test Methods for Constituent Content of Composite Materials, 2000.
[193] L. Thiraphattaraphun, "Structure/Property Relationships in Polypropylene
Nanocomposites," Doctor of Philosophy, School of Materials / Faculty of
Engineering and Physical Sciences, University of Manchester, Manchester 2013.
[194] M. Umar, "Processing, Structure and Properties of PA6/Carbon Composites,"
Doctor of Philosophy (PhD), School of Materials/Faculty of Engineering and Physical
Science, University of Manchester, Manchester, 2014.
[195] Standard Test Method for Tensile Properties of Polymer Matrix Composite Materials
,D 3039/D 3039M – 00, 2006.
[196] Instron. (2018, Online). Non-Contacting Video ExtensometersInstron Available:
http://www.instron.co.uk/en-gb/products/testing-accessories/extensometers/non-
contacting-video/extensometers
[197] ASTM-D790-03 Flexural Properties of Unreinforced and Reinforced Plastics and
Electrical Insulating Materials, 2003.
[198] P. K. Mallick, Fibre-Reinforced composite, 3rd ed. Taylor and Fracis Group, LLC,
2008.

272
[199] A. J. Rodriguez, M. E. Guzman, C.-S. Lim, and B. Minaie, "Mechanical properties of
carbon nanofibre/fibre-reinforced hierarchical polymer composites manufactured
with multiscale-reinforcement fabrics," Carbon, vol. 49, pp. 937-948, 2011.
[200] ASTM D 2344/D 2344M – 00, Standard test method for short-beam strength of
polymer matrix composite materials and their laminates, 2006.
[201] ASTM D 5528 – 01 Standard Test Method for Mode I Interlaminar Fracture
Toughness of Unidirectional Fiber-Reinforced Polymer Matrix Composites, 2002.
[202] D. W. Y. Wong, L. Lin, P. T. McGrail, T. Peijs, and P. J. Hogg, "Improved fracture
toughness of carbon fibre/epoxy composite laminates using dissolvable
thermoplastic fibres," Composites Part A: Applied Science and Manufacturing, vol.
41, no. 6, pp. 759-767, 6// 2010.
[203] M. Hojo, S. Matsuda, M. Tanaka, S. Ochiai, and A. Murakami, "Mode I delamination
fatigue properties of interlayer-toughened CF/epoxy laminates," Composites
Science and Technology, vol. 66, no. 5, pp. 665-675, 2006.
[204] J. C. PRICHARD and P. J. HOGG, "The role of impact damage in post-impact
compression testing," Composites, vol. 21, no. 6, 1990.
[205] ASTM D 7136/D 7136M – 05, Standard Test Method for Measuring the Damage
Resistance of a Fiber-Reinforced Polymer Matrix Composite to a Drop-Weight
Impact Event, 2005.
[206] ASNT, Non-destructive testing handbook, 2nd ed. Ultrasonic testing, American
Society for Non-destructive Testing,, 1991.
[207] M. Dale, B. A. Acha, and L. A. Carlsson, "Low velocity impact and compression after
impact characterization of woven carbon/vinylester at dry and water saturated
conditions," Composite Structures, vol. 94, no. 5, pp. 1582-1589, 2012.
[208] B. A. Alshammari, "Processing, Structure and Properties of Poly (ethylene
terephthalate)/Carbon Micro- and Nano-composites," Doctor of Philosophy (PhD),
Faculty of Engineering and Physical Science, University of Manchester, 2014.
[209] G. Gedler, M. Antunes, V. Realinho, and J. I. Velasco, "Thermal stability of
polycarbonate-graphene nanocomposite foams," Polymer Degradation and
Stability, vol. 97, no. 8, pp. 1297-1304, 2012.
[210] K. Nawaz et al., "Effect of Concentration of Surfactant on the Exfoliation of Graphite
to Graphene in Aqueous Media," Nanomaterials and Nanotechnology, vol. 6, p. 14,
2016.
[211] U. Khan, A. O'Neill, M. Lotya, S. De, and J. N. Coleman, "High-concentration solvent
exfoliation of graphene," Small, vol. 6, no. 7, pp. 864-71, Apr 9 2010.
[212] A. Caradonna, G. Colucci, M. Giorcelli, A. Frache, and C. Badini, "Thermal behavior
of thermoplastic polymer nanocomposites containing graphene nanoplatelets,"
Journal of Applied Polymer Science, vol. 134, no. 20, 2017.
[213] Z. Špitalský, J. Kratochvíla, K. Csomorová, I. Krupa, M. P. F. Graça, and L. C. Costa,
"Mechanical and Electrical Properties of Styrene-Isoprene-Styrene Copolymer
Doped with Expanded Graphite Nanoplatelets," Journal of Nanomaterials, vol.
2015, pp. 1-9, 2015.
[214] S. Dul et al., "Effect of graphene nanoplatelets structure on the properties of
acrylonitrile-butadiene-styrene composites," Polymer Composites, 2017.
[215] F. Wang, L. T. Drzal, Y. Qin, and Z. Huang, "Mechanical properties and thermal
conductivity of graphene nanoplatelet/epoxy composites," Journal of Materials
Science, vol. 50, no. 3, pp. 1082-1093, 2014.
[216] B. D and B. G, High resolution XPS of organic polymers: the scienta esca 300
database. New York: John Wiley and Sons, 1992.

273
[217] B. Finnigan, D. Martin, P. Halley, R. Truss, and K. Campbell, "Morphology and
properties of thermoplastic polyurethane nanocomposites incorporating
hydrophilic layered silicates," Polymer, vol. 45, no. 7, pp. 2249-2260, 2004.
[218] K. N. M. Amin, "Cellulose Nanocrystals Reinforced Thermoplastic Polyurethane
Nanocomposites," degree of Doctor of Philosophy, Australian Institute for
Bioengineering and Nanotechnology (AIBN), University of Queensland, 2016.
[219] M. K. M. Halit, "Proccessing, Structure and Properties of Polyamide 6/Graphene
Nanoplatelets Nanocomposites," Faculty of Science and Engineering, University of
Manchester, 2018.
[220] K. K. Sadasivuni et al., "Dielectric properties of modified graphene oxide filled
polyurethane nanocomposites and its correlation with rheology," Composites
Science and Technology, vol. 104, pp. 18-25, 2014.
[221] P. Pokharel and D. S. Lee, "High performance polyurethane nanocomposite films
prepared from a masterbatch of graphene oxide in polyether polyol," Chemical
Engineering Journal, vol. 253, pp. 356-365, 2014.
[222] J. Zhang, C. Zhang, and S. A. Madbouly, "In situpolymerization of bio-based
thermosetting polyurethane/graphene oxide nanocomposites," Journal of Applied
Polymer Science, vol. 132, no. 13, pp. n/a-n/a, 2015.
[223] A. Yasmin and I. M. Daniel, "Mechanical and thermal properties of graphite
platelet/epoxy composites," Polymer, vol. 45, no. 24, pp. 8211-8219, 2004.
[224] L. Yue, G. Pircheraghi, S. A. Monemian, and I. Manas-Zloczower, "Epoxy composites
with carbon nanotubes and graphene nanoplatelets – Dispersion and synergy
effects," Carbon, vol. 78, pp. 268-278, 2014.
[225] H.-C. Kuan, C.-C. M. Ma, W.-P. Chang, S.-M. Yuen, H.-H. Wu, and T.-M. Lee,
"Synthesis, thermal, mechanical and rheological properties of multiwall carbon
nanotube/waterborne polyurethane nanocomposite," Composites Science and
Technology, vol. 65, no. 11-12, pp. 1703-1710, 2005.
[226] A. Yasmin, J.-J. Luo, and I. M. Daniel, "Processing of expanded graphite reinforced
polymer nanocomposites," Composites Science and Technology, vol. 66, no. 9, pp.
1182-1189, 2006.
[227] S. Bandla and J. C. Hanan, "Microstructure and elastic tensile behavior of
polyethylene terephthalate-exfoliated graphene nanocomposites," J Mater Sci, vol.
47, pp. 876–882, 2012.
[228] T. McNally et al., "Polyethylene multiwalled carbon nanotube composites,"
Polymer, vol. 46, no. 19, pp. 8222-8232, 2005.
[229] B. Ramezanzadeh, E. Ghasemi, M. Mahdavian, E. Changizi, and M. H.
Mohamadzadeh Moghadam, "Characterization of covalently-grafted polyisocyanate
chains onto graphene oxide for polyurethane composites with improved
mechanical properties," Chemical Engineering Journal, vol. 281, pp. 869-883, 2015.
[230] N. Song, J. Yang, P. Ding, S. Tang, and L. Shi, "Effect of polymer modifier chain
length on thermal conductive property of polyamide 6/graphene nanocomposites,"
Composites Part A: Applied Science and Manufacturing, vol. 73, pp. 232-241, 2015.
[231] M. Bera and P. K. Maji, "Effect of structural disparity of graphene-based materials
on thermo-mechanical and surface properties of thermoplastic polyurethane
nanocomposites," Polymer, vol. 119, pp. 118-133, 2017.
[232] R. Shamsi, M. Koosha, and M. Mahyari, "Improving the mechanical, thermal and
electrical properties of polyurethane- graphene oxide nanocomposites synthesized
by in-situ polymerization of ester-based polyol with hexamethylene diisocyanate,"
Journal of Polymer Research, vol. 23, no. 12, 2016.

274
[233] Z. Z. ZORAN S. PETROVlC, JOSEPH H. FLY, and WILLIAM J. MACKNIGHT, "Thermal
Degradation of Segmented Polyurethanes," Applied Polymer Science, vol. 51, no. 6,
pp. 1087-1095, 1994.
[234] M. Kumar, J. S. Chung, B.-S. Kong, E. J. Kim, and S. H. Hur, "Synthesis of graphene–
polyurethane nanocomposite using highly functionalized graphene oxide as
pseudo-crosslinker," Materials Letters, vol. 106, pp. 319-321, 2013.
[235] M. Kotal, T. Kuila, S. K. Srivastava, and A. K. Bhowmick, "Synthesis and
characterization of polyurethane/Mg-Al layered double hydroxide
nanocomposites," Journal of Applied Polymer Science, vol. 114, no. 5, pp. 2691-
2699, 2009.
[236] S. Roy, S. K. Srivastava, J. Pionteck, and V. Mittal, "Mechanically and Thermally
Enhanced Multiwalled Carbon Nanotube-Graphene Hybrid filled Thermoplastic
Polyurethane Nanocomposites," Macromolecular Materials and Engineering, vol.
300, no. 3, pp. 346-357, 2015.
[237] M. Herrera, G. Matuschek, and A. Kettrup, "Thermal degradation of thermoplastic
polyurethane elastomers (TPU) based on MDI," Polymer Degradation and Stability,
vol. 78, pp. 323–331, 2002.
[238] S. Thakur and N. Karak, "A tough, smart elastomeric bio-based hyperbranched
polyurethane nanocomposite," New Journal of Chemistry, vol. 39, no. 3, pp. 2146-
2154, 2015.
[239] S. Thakur and N. Karak, "Ultratough, Ductile, Castor Oil-Based, Hyperbranched,
Polyurethane Nanocomposite Using Functionalized Reduced Graphene Oxide," ACS
Sustainable Chemistry & Engineering, vol. 2, no. 5, pp. 1195-1202, 2014.
[240] E. H. Jeong et al., "Effective preparation and characterization of
montmorillonite/poly(ɛ-caprolactone)-based polyurethane nanocomposites,"
Journal of Applied Polymer Science, vol. 107, no. 2, pp. 803-809, 2008.
[241] A. K. Barick and D. K. Tripathy, "Preparation, characterization and properties of acid
functionalized multi-walled carbon nanotube reinforced thermoplastic
polyurethane nanocomposites," Materials Science and Engineering: B, vol. 176, no.
18, pp. 1435-1447, 2011.
[242] L. Lei, Z. Xia, L. Zhang, Y. Zhang, and L. Zhong, "Preparation and properties of
amino-functional reduced graphene oxide/waterborne polyurethane hybrid
emulsions," Progress in Organic Coatings, vol. 97, pp. 19-27, 2016.
[243] A. Ali, K. Yusoh, and S. F. Hasany, "Synthesis and Physicochemical Behaviour of
Polyurethane-Multiwalled Carbon Nanotubes Nanocomposites Based on
Renewable Castor Oil Polyols," Journal of Nanomaterials, vol. 2014, pp. 1-9, 2014.
[244] A. Frick and A. Rochman, "Characterization of TPU-elastomers by thermal analysis
(DSC)," Polymer Testing, vol. 23, no. 4, pp. 413-417, 2004.
[245] T. Hosseini-Sianaki, H. Nazockdast, B. Salehnia, and E. Nazockdast, "Microphase
separation and hard domain assembly in thermoplastic polyurethane/multiwalled
carbon nanotube nanocomposites," Polymer Engineering & Science, vol. 55, no. 9,
pp. 2163-2173, 2015.
[246] J. Bian, H. L. Lin, F. X. He, X. W. Wei, I. T. Chang, and E. Sancaktar, "Fabrication of
microwave exfoliated graphite oxide reinforced thermoplastic polyurethane
nanocomposites: Effects of filler on morphology, mechanical, thermal and
conductive properties," Composites Part A: Applied Science and Manufacturing, vol.
47, pp. 72-82, 2013.
[247] M. Razeghi and G. Pircheraghi, "TPU/graphene nanocomposites: Effect of graphene
functionality on the morphology of separated hard domains in thermoplastic
polyurethane," Polymer, vol. 148, pp. 169-180, 2018.

275
[248] C. H. Dan, M. H. Lee, Y. D. Kim, B. H. Min, and J. H. Kim, "Effect of clay modifiers on
the morphology and physical properties of thermoplastic polyurethane/clay
nanocomposites," Polymer, vol. 47, no. 19, pp. 6718-6730, 2006.
[249] K. N. Raftopoulos, B. Janowski, L. Apekis, K. Pielichowski, and P. Pissis, "Molecular
mobility and crystallinity in polytetramethylene ether glycol in the bulk and as soft
component in polyurethanes," European Polymer Journal, vol. 47, no. 11, pp. 2120-
2133, 2011.
[250] Y.-C. Chung, N. D. Khiem, and B. C. Chun, "Characterization of a polyurethane
copolymer covalently linked to graphite and the influence of graphite on electric
conductivity," Journal of Composite Materials, vol. 49, no. 14, pp. 1689-1703, 2014.
[251] D. A. Nguyen, Y. R. Lee, A. V. Raghu, H. M. Jeong, C. M. Shin, and B. K. Kim,
"Morphological and physical properties of a thermoplastic polyurethane reinforced
with functionalized graphene sheet," Polymer International, vol. 58, no. 4, pp. 412-
417, 2009.
[252] S. Araby et al., "Melt compounding with graphene to develop functional, high-
performance elastomers," Nanotechnology, vol. 24, no. 16, p. 165601, Apr 26 2013.
[253] M. Karevan and K. Kalaitzidou, "Formation of a complex constrained region at the
graphite nanoplatelets-polyamide 12 interface," Polymer, vol. 54, no. 14, pp. 3691-
3698, 2013.
[254] S. Koutsoumpis, K. N. Raftopoulos, O. Oguz, C. M. Papadakis, Y. Z. Menceloglu, and
P. Pissis, "Dynamic glass transition of the rigid amorphous fraction in polyurethane-
urea/SiO2 nanocomposites," Soft Matter, vol. 13, no. 26, pp. 4580-4590, Jul 14
2017.
[255] T.-L. Wang and C.-G. Tseng, "Polymeric carbon nanocomposites from multiwalled
carbon nanotubes functionalized with segmented polyurethane," Journal of Applied
Polymer Science, vol. 105, no. 3, pp. 1642-1650, 2007.
[256] Y.-Q. Zhang, J.-H. Lee, J. M. Rhee, and K. Y. Rhee, "Polypropylene–clay
nanocomposites prepared by in situ grafting-intercalating in melt," Composites
Science and Technology, vol. 64, no. 9, pp. 1383-1389, 2004.
[257] M. I. Sarwar, S. Zulfiqar, and Z. Ahmad, "Polyamide–silica nanocomposites:
mechanical, morphological and thermomechanical investigations," Polymer
International, vol. 57, no. 2, pp. 292-296, 2008.
[258] A. N. Wilkinson et al., "Structure and Dynamic Mechanical Properties of Melt
Intercalated Polyamide 6—Montmorillonite Nanocomposites," Macromolecular
Materials and Engineering, vol. 291, no. 8, pp. 917-928, 2006.
[259] A. Saritha and K. Joseph, "Effect of nano clay on the constrained polymer volume of
chlorobutyl rubber nanocomposites," Polymer Composites, vol. 36, no. 11, pp.
2135-2139, 2015.
[260] B. Chen, N. Ma, X. Bai, H. Zhang, and Y. Zhang, "Effects of graphene oxide on
surface energy, mechanical, damping and thermal properties of ethylene-
propylene-diene rubber/petroleum resin blends," RSC Advances, vol. 2, no. 11, p.
4683, 2012.
[261] J. C. G. SEEFRIED, J. V. KOLESKE, and F. E. CRITCHFIELD, "Thermoplastic Urethane
Elastomers. I. Effects of Soft -Segment Variations," Applied Polymer Science, vol. 19,
pp. 2493-2502, 1975.
[262] I. A. Kinloch, S. A. Roberts, and A. H. Windle, "A rheological study of concentrated
aqueous nanotube dispersions," Polymer, vol. 43, pp. 7483–7491, 2002.
[263] X. Wang and M. Rackaitis, "Fluctuations and critical phenomena of a filled
elastomer under deformation," Europhysics Letters (EPL), vol. 75, no. 4, pp. 590-
596, 2006.

276
[264] A. K. Barick and D. K. Tripathy, "Nanostructure morphology and dynamic rheological
properties of nanocomposites based on thermoplastic polyurethane and organically
modified montmorillonite," Polymer Bulletin, vol. 66, no. 9, pp. 1231-1253, 2010.
[265] R. Krishnamoorti and K. Yurekli, "Rheology of polymer layered silicate
nanocomposites," Current Opinion in Colloid & Interface Science, vol. 6, p. 464᎐470,
2001.
[266] H. Kim and C. W. Macosko, "Morphology and Properties of Polyester/Exfoliated
Graphite Nanocomposites," Macromolecules, vol. 41, pp. 3317-3327, 2008.
[267] T. Aubry, T. Razafinimaro, and P. Médéric, "Rheological investigation of the melt
state elastic and yield properties of a polyamide-12 layered silicate
nanocomposite," Journal of Rheology, vol. 49, no. 2, pp. 425-440, 2005.
[268] A. K. Barick and D. K. Tripathy, "Effect of organically modified layered silicate
nanoclay on the dynamic viscoelastic properties of thermoplastic polyurethane
nanocomposites," Applied Clay Science, vol. 52, no. 3, pp. 312-321, 2011.
[269] M. J. Kayatin and V. A. Davis, "Viscoelasticity and Shear Stability of Single-Walled
Carbon Nanotube/Unsaturated Polyester Resin Dispersions," Macromolecules, vol.
42, no. 17, pp. 6624-6632, 2009.
[270] Z. Fan and S. G. Advani, "Rheology of multiwall carbon nanotube suspensions,"
Journal of Rheology, vol. 51, no. 4, pp. 585-604, 2007.
[271] J. Sumfleth, S. T. Buschhorn, and K. Schulte, "Comparison of rheological and
electrical percolation phenomena in carbon black and carbon nanotube filled epoxy
polymers," Journal of Materials Science, vol. 46, no. 3, pp. 659-669, 2010.
[272] F. Du, Robert C. Scogna, Wei Zhou, Stijn Brand, John E. Fischer, and Karen I. Winey
"Nanotube networks in polymer nanocomposites: rheology and electrical
conductivity," Macromolecules, vol. 37, pp. 9048-9055, 2004.
[273] M. Annala, M. Lahelin, and J. Seppala, "Utilization of poly(methyl methacrylate) –
carbon nanotube and polystyrene – carbon nanotube in situ polymerized
composites as masterbatches for melt mixing," Express Polymer Letters, vol. 6, no.
10, pp. 814-825, 2012.
[274] X. Zhang et al., "Electrically conductive polypropylene nanocomposites with
negative permittivity at low carbon nanotube loading levels," ACS Appl Mater
Interfaces, vol. 7, no. 11, pp. 6125-38, Mar 25 2015.
[275] X. Yang, Y. Zhan, R. Zhao, and X. Liu, "Effects of graphene nanosheets on the
dielectric, mechanical, thermal properties, and rheological behaviors of
poly(arylene ether nitriles)," Journal of Applied Polymer Science, vol. 124, no. 2, pp.
1723-1730, 2012.
[276] G. L. Hwang, Y. T. Shieh, and K. C. Hwang, "Efficient Load Transfer to Polymer-
Grafted Multiwalled Carbon Nanotubes in Polymer Composites," Advanced
Functional Materials, vol. 14, no. 5, pp. 487-491, 2004.
[277] J.-Z. Liang, Q. Du, G. C.-P. Tsui, and C.-Y. Tang, "Tensile properties of graphene
nano-platelets reinforced polypropylene composites," Composites Part B:
Engineering, vol. 95, pp. 166-171, 2016.
[278] P. Pokharel and D. S. Lee, "Thermal and Mechanical Properties of Reduced
Graphene Oxide/Polyurethane Nanocomposite," Journal of Nanoscience and
Nanotechnology, vol. 14, no. 8, pp. 5718-5721, 2014.
[279] W. J. M. Weiming Tang, and Shaw L. Hsu, "Segmented Polyurethane Elastomers
with Liquid Crystalline Hard Segments. 3. Infrared Spectroscopic Study,"
Macromolecules, vol. 28, pp. 4284-4289, 1995.
[280] M. Liu, C. Zhang, W. W. Tjiu, Z. Yang, W. Wang, and T. Liu, "One-step hybridization
of graphene nanoribbons with carbon nanotubes and its strong-yet-ductile

277
thermoplastic polyurethane composites," Polymer, vol. 54, no. 12, pp. 3124-3130,
2013.
[281] K. Zhou, Z. Gui, Y. Hu, S. Jiang, and G. Tang, "The influence of cobalt oxide–
graphene hybrids on thermal degradation, fire hazards and mechanical properties
of thermoplastic polyurethane composites," Composites Part A: Applied Science and
Manufacturing, vol. 88, pp. 10-18, 2016.
[282] Q. Chen, J. D. Mangadlao, J. Wallat, A. De Leon, J. K. Pokorski, and R. C. Advincula,
"3D Printing Biocompatible Polyurethane/Poly(lactic acid)/Graphene Oxide
Nanocomposites: Anisotropic Properties," ACS Appl Mater Interfaces, vol. 9, no. 4,
pp. 4015-4023, Feb 1 2017.
[283] B. Mayoral et al., "Melt processing and characterisation of polyamide 6/graphene
nanoplatelet composites," RSC Advances, vol. 5, no. 65, pp. 52395-52409, 2015.
[284] J. Liang et al., "Molecular-Level Dispersion of Graphene into Poly(vinyl alcohol) and
Effective Reinforcement of their Nanocomposites," Advanced Functional Materials,
vol. 19, no. 14, pp. 2297-2302, 2009.
[285] D. Han, L. Yan, W. Chen, W. Li, and P. R. Bangal, "Cellulose/graphite oxide
composite films with improved mechanical properties over a wide range of
temperature," Carbohydrate Polymers, vol. 83, no. 2, pp. 966-972, 2011.
[286] U. Khan, P. May, A. O’Neill, and J. N. Coleman, "Development of stiff, strong, yet
tough composites by the addition of solvent exfoliated graphene to polyurethane,"
Carbon, vol. 48, no. 14, pp. 4035-4041, 2010.
[287] L.-C. Tang et al., "The effect of graphene dispersion on the mechanical properties of
graphene/epoxy composites," Carbon, vol. 60, pp. 16-27, 2013.
[288] R. Verdejo, M. M. Bernal, L. J. Romasanta, and M. A. Lopez-Manchado, "Graphene
filled polymer nanocomposites," J. Mater. Chem., vol. 21, no. 10, pp. 3301-3310,
2011.
[289] N. Yousefi, M. M. Gudarzi, Q. Zheng, S. H. Aboutalebi, F. Sharif, and J.-K. Kim, "Self-
alignment and high electrical conductivity of ultralarge graphene oxide–
polyurethane nanocomposites," Journal of Materials Chemistry, vol. 22, no. 25, p.
12709, 2012.
[290] S. D. A. S. Ramôa, G. M. O. Barra, R. V. B. Oliveira, M. G. de Oliveira, M. Cossa, and
B. G. Soares, "Electrical, rheological and electromagnetic interference shielding
properties of thermoplastic polyurethane/carbon nanotube composites," Polymer
International, vol. 62, no. 10, pp. 1477-1484, 2013.
[291] W. Zhenga, Wonga, S.-C., and Sueb, H.-J., "Transport behavior of PMMA/expanded
graphite nanocomposites," Polymer, vol. 73, pp. 6767–6773, 2002.
[292] G. Chen, W. Weng, D. Wu, and C. Wu, "PMMA/graphite nanosheets composite and
its conducting properties," European Polymer Journal, vol. 39, no. 12, pp. 2329-
2335, 2003.
[293] I. Balberg, N. Binenbaum, and N. Wagner, "Percolation Thresholds in the Three-
Dimensional Sticks System," Physical Review Letters, vol. 52, no. 17, pp. 1465-1468,
1984.
[294] S. H. Munson-McGee, "Estimation of the critical concentration in an anisotropic
percolation network," Physical Review B, vol. 43, no. 4, pp. 3331-3336, 1991.
[295] L. Hu, T. Desai, and P. Keblinski, "Thermal transport in graphene-based
nanocomposite," Journal of Applied Physics, vol. 110, no. 3, p. 033517, 2011.
[296] E. Pop, V. Varshney, and A. K. Roy, "Thermal properties of graphene: Fundamentals
and applications," MRS Bulletin, vol. 37, no. 12, pp. 1273-1281, 2012.
[297] Z. Han and A. Fina, "Thermal conductivity of carbon nanotubes and their polymer
nanocomposites: A review," Progress in Polymer Science, vol. 36, no. 7, pp. 914-
944, 2011.

278
[298] S. Y. Pak, H. M. Kim, S. Y. Kim, and J. R. Youn, "Synergistic improvement of thermal
conductivity of thermoplastic composites with mixed boron nitride and multi-
walled carbon nanotube fillers," Carbon, vol. 50, no. 13, pp. 4830-4838, 2012.
[299] D. E. W. Stone and B. Clarke, "Ultrasonic attenuation as a measure of void content
in carbon-fibre reinforced plastics," NON-DESTRUCTIVE TESTING, pp. 137-45, 1975
July
[300] T. Chang, L. Zhan, W. Tan, and S. Li, "Void content and interfacial properties of
composite laminates under different autoclave cure pressure," Composite
interfaCes, vol. 24, no. 5, pp. 529–540, 2017.
[301] P.-O. Hagstrand, F. Bonjour, and J.-A. E. Manson, "The influence of void content on
the structural flexural performance of unidirectional glass fibre reinforced
polypropylene composites," Composites: Part A, vol. 36, pp. 705–714, 2005.
[302] L. Liu, B.-M. Zhang, D.-F. Wang, and Z.-J. Wu, "Effects of cure cycles on void content
and mechanical properties of composite laminates," Composite Structures, vol. 73,
pp. 303–309, 2006.
[303] F. Wang, L. T. Drzal, Y. Qin, and Z. Huang, "Size effect of graphene nanoplatelets on
the morphology and mechanical behavior of glass fiber/epoxy composites," J Mater
Sci, vol. 51, pp. 3337–3348, 2016.
[304] M. H. Gabr, W. Okumura, H. Ueda, W. Kuriyama, K. Uzawa, and I. Kimpara,
"Mechanical and thermal properties of carbon fiber/polypropylene composite filled
with nano-clay," Composites: Part B, vol. 69, pp. 94–100, 2015.
[305] X. Yan, Y. Imai, D. Shimamoto, and Y. Hotta, "Relationship study between crystal
structure and thermal/mechanical properties of polyamide 6 reinforced and
unreinforced by carbon fiber from macro and local view," Polymer, vol. 55, p.
6186e6194, 2014.
[306] S. Ayyagar, M. Al-Haik, and V. Rollin, "Mechanical and Electrical Characterization of
Carbon Fiber/Bucky Paper/Zinc Oxide Hybrid Composites," Journal of Carbon
Research, vol. 4, no. 6, 2018.
[307] D. P. N. Vlasveld, H. E. N. Berseec, and S. J. Picken, "Nanocomposite matrix for
increased fibre composite strength," polymer, vol. 46, pp. 10269–10278, 2005.
[308] Z. Fan, K.-T. Hsiao, and S. G. Advani, "Experimental investigation of dispersion
during flow of multi-walled carbon nanotube/polymer suspension in fibrous porous
media," Carbon, vol. 42, pp. 871–876, 2004.
[309] A. Warrier et al., "The effect of adding carbon nanotubes to glass/epoxy composites
in the fibre sizing and/or the matrix," Composites Part A: Applied Science and
Manufacturing, vol. 41, no. 4, pp. 532-538, 2010/04/01/ 2010.
[310] C. F. Zorowski and T. Murayama, "Wave Propagation and Dynamic Modulus in
Continuous-Filament Twisted Yarns1:Part I: Low Initial Yarn Tension," Textile
Research Journal, vol. 37, no. 10, pp. 852-860, 1967.
[311] Y. Ma et al., "Effects of molecular weight on the dynamic mechanical properties and
interfacial properties of carbon fiber fabric-reinforced polyetherketone cardo
composites," High Performance Polymers, vol. 28, no. 10, pp. 1210–1217, 2016.
[312] X. Yang, Z. Wang, M. Xu, R. Zhao, and X. Liu, "Dramatic mechanical and thermal
increments of thermoplastic composites by multi-scale synergetic reinforcement:
Carbon fiber and graphene nanoplatelet," Materials & Design, vol. 44, pp. 74-80,
2013.
[313] T. Kuilla, S. Bhadra, D. Yao, N. H. Kim, S. Bose, and J. H. Lee, "Recent advances in
graphene based polymer composites," Progress in Polymer Science, vol. 35, no. 11,
pp. 1350-1375, 2010.
[314] J. Kim, B.-s. Yim, J.-m. Kim, and J. Kim, "The effects of functionalized graphene
nanosheets on the thermal and mechanical properties of epoxy composites for

279
anisotropic conductive adhesives (ACAs)," Microelectronics Reliability, vol. 52, no. 3,
pp. 595-602, 2012.
[315] S. Chatterjee et al., "Mechanical reinforcement and thermal conductivity in
expanded graphene nanoplatelets reinforced epoxy composites," Chemical Physics
Letters, vol. 531, pp. 6-10, 2012.
[316] J. Paulo Davim, P. Reis, V. t. Lapa, and C. Conceição António, "Machinability study
on polyetheretherketone (PEEK) unreinforced and reinforced (GF30) for
applications in structural components," Composite Structures, vol. 62, no. 1, pp. 67-
73, 2003.
[317] A. M. Díez-Pascual et al., "High performance PEEK/carbon nanotube composites
compatibilized with polysulfones-II. Mechanical and electrical properties," Carbon,
vol. 48, no. 12, pp. 3500-3511, 2010.
[318] l. Hollaway, Handbook of polymer composites for engineers. Abington Hall.
Abington, Cambridge, England: Woodhead Publishing Ltd, 1994.
[319] M. A. Ghafaar, A. A. Mazen, and N. A. El-Mahallawy, "Application of The Rule of
Mixtures and Halpin-Tsai Equations to Woven Fabric Reinforced Epoxy
Composites," Journal of Engineering Sciences, vol. 34, no. 1, pp. pp. 227-236, 2006.
[320] Z. Li, R. J. Young, N. R. Wilson, I. A. Kinloch, C. Vallés, and Z. Li, "Effect of the
orientation of graphene-based nanoplatelets upon the Young's modulus of
nanocomposites," Composites Science and Technology, vol. 123, pp. 125-133, 2016.
[321] Z. Li, R. J. Young, I. A. Kinloch, N. R. Wilson, A. J. Marsden, and A. P. A. Raju,
"Quantitative determination of the spatial orientation of graphene by polarized
Raman spectroscopy," Carbon, vol. 88, pp. 215-224, 2015.
[322] A. Kelly and T. W.R, "Tensile properties of fibre-reinforced metals: Copper/tungsten
and copper/molybdenum.," Journal of the Mechanics and Physics of Solids, vol. 13,
pp. 329–350, 1965.
[323] S. Syngellakis, Natural Filler and Fibre Composites: Development and
Characterisation. Southampton, UK: WIT Press, 2015.
[324] L. Gong, "Deformation Micromechanics of Graphene Nanocomposites," PhD, School
of Materials University of Manchester, Manchester, UK, 2013.
[325] C. Lee, X. Wei, J. W. Kysa, and J. Hone, "Measurement of the Elastic Properties and
Intrinsic Strength of Monolayer Graphene," Science, vol. 321, pp. 385-388, 2008.
[326] J. A. King, D. R. Klimek, I. Miskioglu, and G. M. Odegard, "Mechanical properties of
graphene nanoplatelet/epoxy composites," Journal of Composite Materials, vol. 49,
no. 6, pp. 659-668, 2014.
[327] M. H and R.-R. F, "Sciences of carbon materials," Universidad de Alicante, San
Vicente del Raspeig Alicante, Spain, 2001.
[328] M. Sánchez, M. Campo, A. Jiménez-Suárez, and A. Ureña, "Effect of the carbon
nanotube functionalization on flexural properties of multiscale carbon fiber/epoxy
composites manufactured by VARIM," Composites Part B: Engineering, vol. 45, no.
1, pp. 1613-1619, 2013.
[329] R. Moriche, M. Sánchez, A. Jiménez-Suárez, S. G. Prolongo, and A. Ureña,
"Electrically conductive functionalized-GNP/epoxy based composites: From
nanocomposite to multiscale glass fibre composite material," Composites Part B:
Engineering, vol. 98, pp. 49-55, 2016.
[330] M. L. Costaa, S. r. F. M. d. Almeidaa, and M. C. Rezendeb, "The influence of porosity
on the interlaminar shear strength of carbon/epoxy and carbon/bismaleimide
fabric laminates," Composites Science and Technology, vol. 61, pp. 2101-2108,
2001.

280
[331] R. Olivier, J. P. Cottu, and B. Ferret, "Effects of cure cycle pressure and voids on
some mechanical properties of carbon/epoxy laminates," Composites, vol. 26, pp.
509-515, 1995.
[332] M. R. Wisnom, T. Reynolds, and N. Gwilliam, "Reduction in Interlaminar Shear
Strength by discrete and distributed voids," Composites Science and Technology vol.
56, pp. 93-101, 1996.
[333] Z. Hong-yan, L. Di-hong, Z. Dong-xing, W. Bao-chang, and C. Yu-yong, "Influence of
voids on interlaminar shear strength of carbon/epoxy fabric laminates,"
Transactions of Nonferrous Metals Society of China, vol. 19, no. s470-s475, 2009.
[334] K. J. Bowles and S. Frimpong. (1991) Relationship between voids and interlaminar
shear strength of polymer matrix composites. International SAMPE Symposium and
Exhibition.
[335] N. De Greef, L. Gorbatikh, S. V. Lomov, and I. Verpoest, "Damage development in
woven carbon fiber/epoxy composites modified with carbon nanotubes under
tension in the bias direction," Composites Part A: Applied Science and
Manufacturing, vol. 42, no. 11, pp. 1635-1644, 2011.
[336] Mohammad Moniruzzaman and K. I. Winey, "Review: Polymer Nanocomposites
Containing Carbon Nanotubes," Macromolecules, vol. 39, pp. 5194-5205, 2006.
[337] N. Yamamoto, R. Guzman de Villoria, and B. L. Wardle, "Electrical and thermal
property enhancement of fiber-reinforced polymer laminate composites through
controlled implementation of multi-walled carbon nanotubes," Composites Science
and Technology, vol. 72, no. 16, pp. 2009-2015, 2012.
[338] H. S. Kim and H. T. Hahn, "Graphite fiber composites interlayered with single-walled
carbon nanotubes," Journal of Composite Materials, vol. 45, no. 10, pp. 1109-1120,
2011.
[339] P.-C. Ma, N. A. Siddiqui, G. Marom, and J.-K. Kim, "Dispersion and functionalization
of carbon nanotubes for polymer-based nanocomposites: A review," Composites
Part A: Applied Science and Manufacturing, vol. 41, no. 10, pp. 1345-1367, 2010.
[340] L. Diaz-Chacon et al., "Graphite Nanoplatelets Composite Materials: Role of the
Epoxy-System in the Thermal Conductivity," Journal of Materials Science and
Chemical Engineering, vol. 03, no. 05, pp. 75-87, 2015.
[341] M. Mehrali et al., "Preparation and characterization of palmitic acid/graphene
nanoplatelets composite with remarkable thermal conductivity as a novel shape-
stabilized phase change material," Applied Thermal Engineering, vol. 61, no. 2, pp.
633-640, 2013.
[342] A. A. Baker, R. Jones, and R. J. Callinan, "Damage Tolerance of Graphite/Epoxy
Composites," Composite Structures, vol. 4, pp. 15-44, 1985.
[343] J. C. Halpin and J. L. Kardos, "The Halpin-Tsai Equations: A Review," Polmer
Engineering and Science, vol. 16, no. 5, 1976.
[344] A. N. Wilkinson et al., "Tensile properties of melt intercalated polyamide 6 –
Montmorillonite nanocomposites," Composites Science and Technology, vol. 67, no.
15-16, pp. 3360-3368, 2007.
[345] Y. Li et al., "In situpolymerization, thermal, damping, and mechanical properties of
multiwalled carbon nanotubes/polyisobutylene-based polyurethane
nanocomposites," Polymer Composites, vol. 36, no. 1, pp. 198-203, 2015.
[346] S. Dul, A. Pegoretti, and L. Fambri, "Effects of the Nanofillers on Physical Properties
of Acrylonitrile-Butadiene-Styrene Nanocomposites: Comparison of Graphene
Nanoplatelets and Multiwall Carbon Nanotubes," Nanomaterials (Basel), vol. 8, no.
9, Aug 29 2018.
[347] T. Gómez-del Río, P. Poza, J. Rodríguez, M. C. García-Gutiérrez, J. J. Hernández, and
T. A. Ezquerra, "Influence of single-walled carbon nanotubes on the effective elastic

281
constants of poly(ethylene terephthalate)," Composites Science and Technology,
vol. 70, no. 2, pp. 284-290, 2010.
[348] J. A. King, D. R. Klimek, I. Miskioglu, and G. M. Odegard, "Mechanical properties of
graphene nanoplatelet/epoxy composites," Journal of Applied Polymer Science, vol.
128, no. 6, pp. 4217-4223, 2013.
[349] P. K. Mallick, Composites Engineering Handbook. New York, NY: Marcel Dekker
1997.
[350] M. van Es, F. Xiqiao, J. van Turnhout, and E. van der Giessen, "Comparing polymer–
clay nanocomposites with conventional composites using composite modeling," in
Specialty polymer additives: principles and applications: CA Malden, MA: Blackwell
Science;, 2001.
[351] K. Kalaitzidou, H. Fukushima, H. Miyagawa, and L. T. Drzal, "Flexural and tensile
moduli of polypropylene nanocomposites and comparison of experimental data to
Halpin-Tsai and Tandon-Weng models," Polymer Engineering & Science, vol. 47, no.
11, pp. 1796-1803, 2007.
[352] S. R. Ahmad, C. Xue, and R. J. Young, "The mechanisms of reinforcement of
polypropylene by graphene nanoplatelets," Materials Science and Engineering: B,
vol. 216, pp. 2-9, 2017.
[353] L. Gong, R. J. Young, I. A. Kinloch, I. Riaz, R. Jalil, and K. S. Novoselov, "Optimizing
the Reinforcement of Polymer-Based Nanocomposites by Graphene," ACS Nano,
vol. 6, pp. 2086-2095, 2012.
[354] A. J. Duguay, J. W. Nader, A. Kiziltas, D. J. Gardner, and H. J. Dagher, "Exfoliated
graphite nanoplatelet-filled impact modified polypropylene nanocomposites:
influence of particle diameter, filler loading, and coupling agent on the mechanical
properties," Applied Nanoscience, vol. 4, no. 3, pp. 279-291, 2013.
[355] D. L. Hunston, R. J. Moulton, N. J. Johnston, and W. Bascom, "Matrix Resin Effects in
Composite Delamination: Mode I Fracture Aspects," Toughened Composites,
ASTMSTP 937, pp. 74–94, 1987. American Society for Testing and Materials,
Philadelphia
[356] M. Kuwata and P. J. Hogg, "Interlaminar toughness of interleaved CFRP using non-
woven veils: Part 1. Mode-I testing," Composites Part A: Applied Science and
Manufacturing, vol. 42, no. 10, pp. 1551-1559, 2011.
[357] J.-K. Kim and Y.-W. Mai, "High Strength, High Fracture Toughness Fibre Composites
with Interface Control A Review," Composites Science and Technology vol. 41, pp.
333-378, 1991.
[358] L. M. A.Godara, F. Luizi;A. Warrier, S.V. Lomov, A.W. Vuure, L. Gorbatikh,
P.Moldenaers, I.Verpoest, "Influence of carbon nanotube reinforcement on the
processing and the mechanical behaviour of carbon fibre/epoxy composite,"
Carbon, vol. 47, pp. 2914-2923, 2009.
[359] M.-Y. Shen et al., "Mechanical Properties and Tensile Fatigue of Graphene
Nanoplatelets Reinforced Polymer Nanocomposites," Journal of Nanomaterials, vol.
2013, pp. 1-9, 2013.
[360] W. D. BASCOM, D. J. BOLL, B. FULLER, and P. J. PHILLIPS, "Fractography of the
interlaminar fracture of carbon-fibre epoxy composites," Journal of Materials
Science, vol. 20, pp. 3184 - 3190, 1985.
[361] L. A. L. Franco, M. L. A. Graça, and F. S. Silva, "Fractography analysis and fatigue of
thermoplastic composite laminates at different environmental conditions,"
Materials Science and Engineering: A, vol. 488, no. 1-2, pp. 505-513, 2008.
[362] E. S. Greenhalgh, C. Rogers, and P. Robinson, "Fractographic observations on
delamination growth and the subsequent migration through the laminate,"
Composites Science and Technology, vol. 69, no. 14, pp. 2345-2351, 2009.

282
[363] D. PURSLOW, "Matrix fractography of fibrereinforced epoxy composites,"
Composites, vol. 17, 1986.
[364] J. Bonhomme, A. Argüelles, J. Viña, and I. Viña, "Fractography and failure
mechanisms in static mode I and mode II delamination testing of unidirectional
carbon reinforced composites," Polymer Testing, vol. 28, no. 6, pp. 612-617, 2009.
[365] B. Ashrafi et al., "Enhancement of mechanical performance of epoxy/carbon fiber
laminate composites using single-walled carbon nanotubes," Composites Science
and Technology, vol. 71, no. 13, pp. 1569-1578, 2011.
[366] C. Atas and O. Sayman, "An overall view on impact response of woven fabric
composite plates," Composite Structures, vol. 82, no. 3, pp. 336-345, 2008.
[367] S.-L. Gao and J.-K. Kim, "Cooling rate influences in carbon fibre/PEEK composites.
Part III: impact damage performance," Composites: Part A, vol. 32, pp. 775-785,
2001.
[368] M. W. Wardle and G. E. Zahr, S. L. Kessler, G. C. Adams, S. B. Driscoll, and D. R.
Ireland, Eds. Instrumented Impact Testing of Aramid-Reinforced Composite
Materials ( Instrumented Impact Testing of Plastics and Composites Materials,
ASTM STP 936). Philadelphia, USA,: American Society for Testing and Materials,,
1987.
[369] M. N. G. Nejhad and A. Parvizi-Majidi, "Impact behaviour and damage tolerance of
woven carbon fibrereinforced thermoplastic composites," Composites, vol. 21, no.
2, pp. 155-168, 1990.
[370] M. V. Hosur, F. Chowdhury, and S. Jeelani, "Low-Velocity Impact Response and
Ultrasonic NDE of Woven Carbon/ Epoxy—Nanoclay Nanocomposites," Journal of
Composite Materials, vol. 41, no. 18, pp. 2195-2212, 2016.
[371] Z. Asaee, M. Mohamed, D. D. Cicco, and F. Taheri, "Low-Velocity Impact Response
and Damage Mechanism of 3D Fiber-Metal Laminates Reinforced with Amino-
Functionalized Graphene Nanoplatelets," International Journal of Composite
Materials, vol. 7, no. 1, pp. 20-36, 2017.
[372] M. V. Hosur, M. Abdullah, and S. Jeelani, "Manufacturing and low-velocity impact
characterization of foam filled 3-D integrated core sandwich composites with
hybrid face sheets," Composite Structures, vol. 69, no. 2, pp. 167-181, 2005.
[373] M. Sayer, N. B. Bektaş, E. Demir, and H. Çallioğlu, "The effect of temperatures on
hybrid composite laminates under impact loading," Composites Part B: Engineering,
vol. 43, no. 5, pp. 2152-2160, 2012.
[374] I. Taraghi, A. Fereidoon, and F. Taheri-Behrooz, "Low-velocity impact response of
woven Kevlar/epoxy laminated composites reinforced with multi-walled carbon
nanotubes at ambient and low temperatures," Materials & Design, vol. 53, pp. 152-
158, 2014.
[375] A. P. Mouritz, "Review of z-pinned composite laminates," Composites Part A:
Applied Science and Manufacturing, vol. 38, no. 12, pp. 2383-2397, 2007.
[376] H. K. M. Dalfi, "Damage Tolerance of Composite Laminate Toughened via Yarn-level
Hybridisation and New Fibre Architecture," Doctor of Philosophy Faculty of Science
and Engineering University of Manchester, Manchester, UK, 2018.

283
Appendix A

Table A.1: Tg and change in heat capacity (Cp) of hard phase of TPU-70 HS and
TPU70-NCs incorporating with 5 wt.% GNPs M5, GNPsM15 and GNPM25
recorded during first heating DSC scan before annealing
Sample Non-annealed
TgHP (C) Cp (J/g.C)
TPU-70 HS 47.6  0.5 0.075  0.029
TPU70/5wt%GNPM5-NCs 55.2  0.1 0.055  0.043
TPU70/5wt%GNPM15-NCs 56.2  0.3 0.028  0.001
TPU70/5wt%GNPM25-NCs 57.7  1.2 0.085  0.015
Table A.2: Tg and change in heat capacity (Cp) of hard phase of TPU-70 HS and
TPU70-NCs incorporating with 5 wt.% GNPs M5, GNPsM15 and GNPM25
recorded during first heating DSC scan after annealing
Sample Annealed
TgHP (C) Cp (J/g.C)
TPU-70 HS 45.9  0.4 0.017  0.013
TPU70/5wt%GNPM5-NCs 49.4  0.4 0.091  0.006
TPU70/5wt%GNPM15-NCs 48.9  0.3 0.079  0.004
TPU70/5wt%GNPM25-NCs 50.6  2.1 0.059  0.005

Table A.3: The data of Storage modulus of E’ of neat TPU-70 HS and TPU70-NCs
incorporating with 5 wt.% of various GNPs sizes (5, 15 and 25 m denoted by
GNPM5, GNPM15 and GNPM25 respectively) obtained from DMTA test before and
after annealing at three different temperature
Sample Non-Annealed
E’ (MPa) E’ (MPa) E’ (MPa)
-100 C 25 C 100 C
TPU-70 HS 2717  81 534  16 73.3  2.4
TPU70/5wt%GNPM5-NCs 4846  383 1361  84 209  8
TPU70/5wt%GNPM15-NCs 3709  327 1150  17 168  7
TPU70/5wt%GNPM25-NCs 3755  297 1212  72 165  11
Annealed
Sample E’ (MPa) E’ (MPa) E’ (MPa)
-100 C 25 C 100 C
TPU-70 HS 2105  190 550.0  6.5 100.6  8.2
TPU70/5wt%GNPM5/NCs 4803  186 1079.0  57.5 207.0  2.15
TPU70/5wt%GNPM15/NCs 3993  220 1106.0  73.2 296.0  15.6
TPU70/5wt%GNPM25/NCs 3096  30 1156.0  23.5 256.0  4.3

284
Table A.4: Tabulated experimental data of tensile test for TPU-70 HS and
TPU70/GNPs-NCs based on the various size of GNPs(5,15 and 25 m denoted by
M5,M15 and M25 respectively) shows the values of elastic modulus, E, yeild stress,
y, tensile strength, u, and elongation at break %,u
Sample Non-annealed Annealed
E (GPa) E(GPa)
TPU-70 HS 0.20 ± 0.02 0.32 ± 0.02
TPU70/5wt%GNPM5-NCs 0.28 ± 0.05 0.31 ± 0.09
TPU70/5wt%GNPM15-NCs 0.55 ± 0.10 0.54 ± 0.04
TPU70/5wt%GNPM25-NCs 0.74 ± 0.17 0.54 ± 0.06
Samples Yeild stress, y (MPa) Yeild stress, y (MPa)
TPU-70 HS 9.79 ± 0.79 17.25 ± 0.75
TPU70/5wt%GNPM5-NCs 12.32 ± 1.13 20.32 ± 1.46
TPU70/5wt%GNPM15-NCs 19.41 ± 1.90 18.67 ± 1.03
TPU70/5wt%GNPM25-NCs 27.51 ± 0.82 22.24 ± 1.06
Samples Tensile strength, u Tensile strength, u
(MPa) (MPa)
TPU-70 HS 26.15 ± 1.25 34.36 ± 2.67
TPU70/5wt%GNPM5-NCs 20.58 ± 0.97 25.00 ± 0.37
TPU70/5wt%GNPM15-NCs 23.52 ± 0.93 24.95 ± 0.52
TPU70/5wt%GNPM25-NCs 26.29 ± 0.90 25.44 ± 0.38
Samples Elongation at StDv Elongation at StDv
break,u, % break,u, %
TPU-70 HS 299 ± 55.3 249 ± 49
TPU70/5wt%GNPM5/NCs 172.1 ± 51.1 64.1± 26.1
TPU70/5wt%GNPM15/NCs 223.4 ± 18.2 102.1± 6.8
TPU70/5wt%GNPM25/NCs 194.2 ± 14.6 109.5 ± 15.9

285
Table A.5: Tabulated experimental data of flexural test for TPU-70 HS and
TPU70/GNPs-NCs based on the various size of GNPs(5,15 and 25 m denoted by
M5,M15 and M25 respectively) shows the values of flexural strength, f, flexural
modulus Ef, and flexural strain %, f

Samples Non-Annealed Annealed


Flexural Strength, f, Flexural Strength, f,
(MPa) (MPa)
TPU-70 HS 37.07 ± 1.01 39.36 ± 0.66
TPU70/5wt%GNPM5-NCs 39.73 ± 0.15 42.87 ± 0.27
TPU70/5wt%GNPM15-NCs 43.09 ± 0.44 44.66 ± 0.40
TPU70/5wt%GNPM25-NCs 42.43 ± 0.43 40.13 ± 0.56
Samples Flexural Modulus, Ef, Flexural Modulus, Ef,
(GPa) (GPa)
TPU-70 HS 0.76 ± 0.02 0.79 ± 0.03
TPU70/5wt%GNPM5-NCs 1.13 ± 0.03 1.32 ± 0.01
TPU70/5wt%GNPM15-NCs 1.21 ± 0.08 1.29 ± 0.07
TPU70/5wt%GNPM25-NCs 1.18 ± 0.02 1.10 ± 0.01
Samples Flexural Strain, f, % Flexural strain, f,
(GPa)
TPU-70 HS 9.05 ± 0.55 8.7 ± 0.2
TPU70/5wt%GNPM5-NCs 7.62 ± 0.25 7.5 ± 0.6
TPU70/5wt%GNPM15-NCs 8.24 ± 0.25 6.1 ± 0.75
TPU70/5wt%GNPM25-NCs 7.04 ± 0.03 6.03 ± 0.35

286
Appendix B

Table B.1: Actual and normalised average values of flexural strength (FS) and
flexural modulus (FM) of TPU70/CF composites and MSCs based on the different
average diameter sizes of GNPs (5, 15 and 25 m) before annealing. The normalised
values have been calculated according to the maximum Vf = 52%
Sample Average FS Normalised Actual Normalised
(MPa)  average FS average average FM
StDv (MPa)  StDv FM (GPa) (GPa)  StDv
 StDv
TPU70/CF 122.3  9.1 235.3  17.4 14.5  1.3 27.9  2.5
TPU70/5wt.%GNPM5/CF 131.6  4.4 253.1  8.4 19.2  1.5 37.3  2.8
TPU70/5wt.%GNPM15/CF 136.3  6.1 262.2  11.8 21.5  4.1 41.4  7.9
TPU70/5wt.%GNPM25/CF 153.6  6.9 295.3  13.4 29.8  1.9 57.3  3.8

Table B.2: Actual and normalised average values of flexural strength (FS) and
flexural modulus (FM) of TPU70/CF composites and MSCs based on the different
average diameter sizes of GNPs (5, 15 and 25 m) after annealing. The normalised
values have been calculated according to the maximum Vf = 52%
Sample Average FS Normalised Actual Normalised
(MPa) average FS average average FM
(MPa) FM (GPa) (GPa)
TPU70/CF 132.6  4.0 254.8  7.7 19.3  2.3 37.1  4.4
TPU70/5wt.%GNPM5/CF 144.3  9.8 277.5  18.8 24.0  7.1 46.2  7.8
TPU70/5wt.%GNPM15/CF 157.4  6.9 302.8  13.4 26.9  2.2 51.8  4.2
PU70/5wt.%GNPM25/CF 146.5  7.8 281.8  14.5 28.7  3.1 55.1  5.9

Table B. 3: Average and normalised values of interlaminar shear strength (ILSS) of


TPU70/CF laminates and MSCs based on different sizes of GNPs (5, 15 and 25 m)
before and after annealing. The normalised values have been calculated according
to the maximum Vf = 47%.
Non- Annealed Non-annealed Annealed
Sample
annealed
Av. max Av. max Av. Av.
ILSS ILSS normalised normalised
(MPa) (MPa) ILSS (MPa) ILSS (MPa)
TPU70/CF 19.6  0.4 18.2  0.6 9.8  0.48 9.1  0.57
TPU70/5wt.%GNPM5/CF 12.6  0.5 11.4  0.7 6.4  0.46 5.8  0.68
TPU70/5wt.%GNPM15/CF 12.9  0.2 12.3  0.14 6.5  0.1 6.3  0.2
TPU70/5wt.%GNPM25/CF 17.3  0.2 16.8  0.2 8.8  0.25 8.4  0.21

287

You might also like